This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

thanks: This research was supported by by NFSC, grant 12201381.

A relative Nadel-type vanishing theorem

Jingcao Wu [email protected] School of Mathematica, Shanghai University of Finance and Economics, Shanghai 200433, People’s Republic of China
Abstract

In this paper we prove a relative Nadel-type vanishing theorem on Kähler morphisms. Then we discuss its applications on harmonic bundles. In particular, it gives a relative vanishing concerning Saito’s S-sheaves.

:
32J25 (primary), 32L20 (secondary).
keywords:
vanishing theorem, invariance of plurigenera, asymptotic multiplier ideal sheaf.

1 Introduction

The Nadel vanishing theorem is a powerful tool in complex geometry. It says that

Theorem 1.1 ((c.f. [Nad90, Dem93])).

Let XX be a projective manifold of dimension nn, and let LL be a holomorphic line bundle on XX. Fix a Kähler metric ω\omega on XX. Assume that LL is equipped with a singular Hermitian metric φ\varphi with iΘL,φεωi\Theta_{L,\varphi}\geqslant\varepsilon\omega for some ε>0\varepsilon>0. Then

Hq(X,KXL(φ))=0H^{q}(X,K_{X}\otimes L\otimes\mathscr{I}(\varphi))=0

for q>0q>0.

Here (φ)\mathscr{I}(\varphi) refers to the multiplier ideal sheaf [Nad90] associated to φ\varphi. Note that in Theorem 1.1, LL is by definition a big line bundle [Dem12]. So we will also directly say that (L,φ)(L,\varphi) is a big line bundle.

However, one of the limitations of Theorem 1.1 is that LL should be big. It is then asked to generalise it. One direction is to consider the line bundle LL possessing weaker positivity, which is fully studied by [Cao14, Eno93, Mat14, Mat15a, Mat15b, Mat18, Wu17, Wu22]. Another direction should be the relative case. See [Laz04b], Generalisations 9.1.22 and 11.2.15, for example. In this paper, we will provide a relative version of the Nadel-type vanishing theorem. Before stating the main result, we should fix some notations and conventions first.

Throughout this paper, unless otherwise stated, f:XYf:X\rightarrow Y is a proper, locally Kähler morphism from a complex manifold XX to a reduced, irreducible analytic space YY. (Remember that ff is locally Kähler if f1(U)f^{-1}(U) is a Kähler space for any relatively compact, open subset UU of YY.) Every connected component of XX is mapped surjectively to YY. XoX^{o} is a Zariski open subset of XX, and (E,h)(E,h) is a tame, Nakano semi-positive vector bundle over XoX^{o}. Let i:XoXi:X^{o}\rightarrow X be the natural embedding, and define

E(h)x:={s(iE)x;|s|h is locally L2-integrable against the Lebesgue measure}.E(h)_{x}:=\{s\in(i_{\ast}E)_{x};|s|_{h}\textrm{ is locally }L^{2}\textrm{-integrable against the Lebesgue measure}\}.

It is a coherent sheaf on XX by [ShZ21a], Proposition 2.9. In the end, we denote by ll the dimension of a general fibre of ff. Then the main theorem is as follows.

Theorem 1.2.

Let LL be a holomorphic line bundle on XX with κ(L,f)0\kappa(L,f)\geqslant 0. Assume that there exists a section ss of some multiple LmL^{m} such that {s=0}Xo\{s=0\}\subseteq X^{o}. Then

Rqf(KXLE(h)(f,L))=0R^{q}f_{\ast}(K_{X}\otimes L\otimes E(h)\otimes\mathscr{I}(f,\|L\|))=0

for q>lκ(L,f)q>l-\kappa(L,f).

Here (f,L)\mathscr{I}(f,\|L\|) is the relative asymptotic multiplier ideal sheaf and κ(L,f)\kappa(L,f) is the relative Iitaka dimension of LL. Note that when Xo=XX^{o}=X, EE is always tame and E(h)=EE(h)=E. At this time our result extends Generlisation 11.2.15 in [Laz04b].

Corollary 1.1.

Let f:XYf:X\rightarrow Y be a proper, locally Kähler morphism from a complex manifold XX to a reduced, irreducible analytic space YY. Every connected component of XX is mapped surjectively to YY. Let EE be a Nakano semi-positive vector bundle over XX. Let LL be a holomorphic line bundle on XX with κ(L,f)0\kappa(L,f)\geqslant 0. Then

Rqf(KXLE(f,L))=0R^{q}f_{\ast}(K_{X}\otimes L\otimes E\otimes\mathscr{I}(f,\|L\|))=0

for q>lκ(L,f)q>l-\kappa(L,f).

In general, as is explained in [ScY20], EE can be interpreted as a vector bundle over XX equipped with a singular Hermitian metric hh, and E(h)E(h) is a higher rank analogy of a singular Hermitian line bundle tensoring with its associated multiplier ideal sheaf. Moreover, the reflexivity of XoX^{o} allows fruitful applications, and the one on harmonic bundles is extremely interesting among them.

Remember that the harmonic bundles are important objects in non-abelian Hodge theory. More precisely, C. Simpson [Sim92] used it to establish a correspondence between local systems and semistable Higgs bundles with vanishing Chern classes. Then our vanishing implies

Corollary 1.2.

Let LL be a holomorphic line bundle with κ(L,f)0\kappa(L,f)\geqslant 0. Assume that there exists a section ss of some multiple LmL^{m} such that {s=0}Xo\{s=0\}\subseteq X^{o}. Let (H,θ,h)(H,\theta,h) be a tame harmonic bundle over XoX^{o}, and let EE be a subbundle of HH with vanishing second fundamental form and θ¯(E)=0\bar{\theta}(E)=0. Then

Rqf(KXLE(h)(f,L))=0R^{q}f_{\ast}(K_{X}\otimes L\otimes E(h)\otimes\mathscr{I}(f,\|L\|))=0

for q>lκ(L,f)q>l-\kappa(L,f).

The SS-sheaf is a typical example that satisfies the conditions in Corollary 1.2. Recall that M. Saito constructed a coherent sheaf S(ICX(𝕍))S(\mathrm{IC}_{X}(\mathbb{V})) in [Sai91], which is called SS-sheaf, for a real variation of polarized Hodge structure (𝕍,,,S)(\mathbb{V},\nabla,\mathcal{F}^{\cdot},S) [CKS86] on XoX^{o}. It plays a key role to solve Kollar’s conjecture [Ko86b]. For this sheaf, we have

Corollary 1.3.

Let LL be a holomorphic line bundle on XX with κ(L,f)0\kappa(L,f)\geqslant 0. Assume that there exists a section ss of some multiple LmL^{m} such that {s=0}Xo\{s=0\}\subseteq X^{o}. Then

Rqf(S(ICX(𝕍))L(f,L))=0R^{q}f_{\ast}(S(\mathrm{IC}_{X}(\mathbb{V}))\otimes L\otimes\mathscr{I}(f,\|L\|))=0

for q>lκ(L,f)q>l-\kappa(L,f).

It can also be viewed as a refinement of the torsion-freeness part in Kollar’s conjecture, and a generalisation of [ShZ21b], Theorem 1.2.

Now we outline the strategy to prove Theorem 1.2. Firstly, we prove the following Kollár-type injectivity and torsion-free theorem.

Theorem 1.3.

Let LL be a holomorphic line bundle on XX with κ(L,f)0\kappa(L,f)\geqslant 0. For a (non-zero) section ss of some multiple Lm1L^{m-1} such that {s=0}Xo\{s=0\}\subseteq X^{o}, the multiplication map induced by the tensor product with ss

Φ:Rqf(KXLE(h)(f,L))Rqf(KXLmE(h)(f,Lm))\Phi:R^{q}f_{\ast}(K_{X}\otimes L\otimes E(h)\otimes\mathscr{I}(f,\|L\|))\rightarrow R^{q}f_{\ast}(K_{X}\otimes L^{m}\otimes E(h)\otimes\mathscr{I}(f,\|L^{m}\|))

is well-defined and injective for any q0q\geqslant 0. In particular, Rqf(KXLE(h)(f,L))R^{q}f_{\ast}(K_{X}\otimes L\otimes E(h)\otimes\mathscr{I}(f,\|L\|)) is torsion-free for every qq.

The injectivity theorem has been fully studied in [Eno93, Fuj12, Ko86a, Ko86b, Mat14, Mat15a, Mat15b, Mat18, ShZ21a]. However, we cannot directly apply any result among these papers to obtain Theorem 1.3. The reason is that in general there does not exist a metric φ\varphi on LL such that

iΘL,φ0and(φ)=(f,L).i\Theta_{L,\varphi}\geqslant 0\quad\textrm{and}\quad\mathscr{I}(\varphi)=\mathscr{I}(f,\|L\|).

We will follow the idea of [Tak95] to develop a harmonic theory concerning the singular Hermitian metric, and use it to finish the proof.

Note that the injectivity and torsion-free theorem (i.e. Theorem 1.3) together with the Kollár-type vanishing theorem and decomposition theorem is usually called Kollár’s package. In particular, the injectivity and torsion-free theorem is the key among them. In other words,

KXLE(h)(f,L))K_{X}\otimes L\otimes E(h)\otimes\mathscr{I}(f,\|L\|))

should moreover satisfy Kollár’s package. We will leave this work in the future.

Return to our main result, Theorem 1.2 is now a combination of Theorem 1.3 and the Nadel-type vanishing theorem (Theorem 1.3) in [Wu22].

This paper is organised as follows. We first recall some background materials, including the asymptotic multiplier ideal sheaf, tame vector bundles and so on. Then, we proceed to develop the harmonic theory associated with singular Hermitian metrics in Sect.3. Based on this theory, we will prove Theorems 1.2 and 1.3 in Sect.4. In the end, we discuss the applications.

Finally, I would like to mention a series of brilliant work concerning Kollár’s package due to Junchao Shentu and Chen Zhao, including [ShZ21a, ShZ21b, ShZ22]. After a very early prototype of Theorem 1.2 was finished, the author fortunately learned these papers, which inspired the further study of the applications of our vanishing theorem. So I own a lot to them.

2 Preliminary

In this section we will introduce some basic materials.

2.1 The asymptotic multiplier ideal sheaf

This part is mostly collected from [Laz04b].

First recall the definition of the multiplier ideal sheaf associated to an ideal sheaf 𝔞𝒪X\mathfrak{a}\subset\mathcal{O}_{X} and a positive real number cc. Let μ:X~X\mu:\tilde{X}\rightarrow X be a smooth modification such that μ𝔞=𝒪X~(E)\mu^{\ast}\mathfrak{a}=\mathcal{O}_{\tilde{X}}(-E), where EE has the simple normal crossing support. Then the multiplier ideal sheaf is defined as

(c𝔞):=μ𝒪X~(KX~/XcE).\mathscr{I}(c\cdot\mathfrak{a}):=\mu_{\ast}\mathcal{O}_{\tilde{X}}(K_{\tilde{X}/X}-\lfloor cE\rfloor).

Here KX~/XK_{\tilde{X}/X} is the relative canonical bundle and E\lfloor E\rfloor means the round-down.

Now let f:XYf:X\rightarrow Y be a proper, locally Kähler and surjective morphism from a complex manifold XX to a reduced, irreducible analyitc space YY. Suppose that LL is a line bundle on XX whose restriction to a general fibre of ff has non-negative Iitaka dimension. For a positive integer kk, there is a naturally defined homomorphism

ρk:ff(Lk)Lk.\rho_{k}:f^{\ast}f_{\ast}(L^{k})\rightarrow L^{k}.

The relative base-ideal 𝔞k,f\mathfrak{a}_{k,f} of |Lk||L^{k}| is then defined as the image of the induced homomorphism

ffLkLk𝒪X.f^{\ast}f_{\ast}L^{k}\otimes L^{-k}\rightarrow\mathcal{O}_{X}.

Hence for a given positive real number cc, we have the multiplier ideal sheaf (ck𝔞k,f)\mathscr{I}(\frac{c}{k}\cdot\mathfrak{a}_{k,f}) which is also denoted by (f,ck|Lk|)\mathscr{I}(f,\frac{c}{k}|L^{k}|). It is not hard to verify that for every integer p1p\geqslant 1 one has the inclusion

(f,ck|Lk|)(f,cpk|Lpk|).\mathscr{I}(f,\frac{c}{k}|L^{k}|)\subseteq\mathscr{I}(f,\frac{c}{pk}|L^{pk}|).

Therefore the family of ideals

{(f,ck|Lk|)}(k0)\{\mathscr{I}(f,\frac{c}{k}|L^{k}|)\}_{(k\geqslant 0)}

has a unique maximal element from the ascending chain condition on ideals.

Definition 2.1.

The relative asymptotic multiplier ideal sheaf associated to ff, cc and |L||L|,

(f,cL)\mathscr{I}(f,c\|L\|)

is defined to be the unique maximal member among the family of ideals {(f,ck|Lk|)}\{\mathscr{I}(f,\frac{c}{k}|L^{k}|)\}.

Next, we explain the analytic counterpart of the relative multiple ideal sheaf. By definition,

(f,cL)=(f,ck|Lk|)=(ck𝔞k,f)\mathscr{I}\bigl{(}f,c\|L\|\bigr{)}=\mathscr{I}\bigl{(}f,\frac{c}{k}|L^{k}|\bigr{)}=\mathscr{I}\bigl{(}\frac{c}{k}\cdot\mathfrak{a}_{k,f}\bigr{)}

for some kk. In this case, we will say that kk computes (f,cL)\mathscr{I}(f,c\|L\|). Let UU be a Stein open subset of YY. By definition, we can pick {u1,,um}\{u_{1},...,u_{m}\} in Γ(f1(U),Lk)\Gamma(f^{-1}(U),L^{k}) which generate 𝔞k,f\mathfrak{a}_{k,f} on f1(U)f^{-1}(U). Let φU=1klog(|u1|2++|um|2)\varphi_{U}=\frac{1}{k}\log(|u_{1}|^{2}+\cdots+|u_{m}|^{2}) which is a singular metric on L|f1(U)L|_{f^{-1}(U)}. We verify that

(f,ck|Lk|)=(cφU) on f1(U).\mathscr{I}(f,\frac{c}{k}|L^{k}|)=\mathscr{I}(c\varphi_{U})\textrm{ on }f^{-1}(U).

Indeed, let μ:X~X\mu:\tilde{X}\rightarrow X be a smooth modification of 𝔞k,f\mathfrak{a}_{k,f}. Then μ𝔞k,f=𝒪X~(E)\mu^{\ast}\mathfrak{a}_{k,f}=\mathcal{O}_{\tilde{X}}(-E) such that E+except(μ)E+\textrm{except}(\mu) has the simple normal crossing support. Here except(μ)\textrm{except}(\mu) is the exceptional divisor of μ\mu. Now it is computed in [Dem12] that

(cφU)=μ𝒪X~(KX~/XckE) on f1(U)\mathscr{I}\bigl{(}c\varphi_{U}\bigr{)}=\mu_{\ast}\mathcal{O}_{\tilde{X}}\bigl{(}K_{\tilde{X}/X}-\lfloor\frac{c}{k}E\rfloor\bigr{)}\textrm{ on }f^{-1}(U)

which coincides with the definition of (f,ck|Lk|)\mathscr{I}(f,\frac{c}{k}|L^{k}|). Furthermore, if v1,,vmv_{1},...,v_{m} are alternative generators and ψU=1klog(|v1|2++|vm|2)\psi_{U}=\frac{1}{k}\log(|v_{1}|^{2}+\cdots+|v_{m}|^{2}), obviously we have (cφU)=(cψU)\mathscr{I}(c\varphi_{U})=\mathscr{I}(c\psi_{U}). Hence all the (cφU)\mathscr{I}(c\varphi_{U}) patch together to give a globally defined multiplier ideal sheaf (cφ)\mathscr{I}(c\varphi) such that

(cφ)=(f,ck|Lk|)=(f,cL).\mathscr{I}(c\varphi)=\mathscr{I}(f,\frac{c}{k}|L^{k}|)=\mathscr{I}(f,c\|L\|).

Note that {f1(U),φU}\{f^{-1}(U),\varphi_{U}\} does not give a globally defined metric on LL in general. The φ\varphi is interpreted as the collection of functions {f1(U),φU}\{f^{-1}(U),\varphi_{U}\} by abusing the notation, which is called the collection of (local) singular metrics on LL associated to (f,cL)\mathscr{I}(f,c\|L\|). Certainly it depends on the choice of kk and is not unique.

The following elementary property is due to [Laz04b].

Proposition 2.1.

Let L1,L2L_{1},L_{2} be holomorphic line bundles on XX with κ(L1,f),κ(L2,f)0\kappa(L_{1},f),\kappa(L_{2},f)\geqslant 0. Let mm and kk be non-negative integers. Let 𝔞m,f\mathfrak{a}_{m,f} be the base-ideal of |L1m||L^{m}_{1}| relative to ff. Then

𝔞m,f(f,L2k)(f,L1mL2k).\mathfrak{a}_{m,f}\cdot\mathscr{I}(f,\|L^{k}_{2}\|)\subseteq\mathscr{I}(f,\|L^{m}_{1}\otimes L^{k}_{2}\|).

2.2 Tame bundle

Let XoX^{o} be a Zariski open subset of XX, and let (E,h)(E,h) be a Hermitian vector bundle over XoX^{o}.

Definition 2.2.

(E,h)(E,h) is tame on XX, if for every xXx\in X there exist an open neighbourhood UU containing xx, a proper bimeromorphic morphism π:U~U\pi:\tilde{U}\rightarrow U which is biholomorphic on UXoU\cap X^{o}, and a Hermitian vector bundle (Q,hQ)(Q,h_{Q}) on U~\tilde{U} such that

  1. (1)

    πE|UXoQ|π1(UXo)\pi^{\ast}E|_{U\cap X^{o}}\subseteq Q|_{\pi^{-1}(U\cap X^{o})};

  2. (2)

    There is a Hermitian metric hQh^{\prime}_{Q} on Q|π1(UXo)Q|_{\pi^{-1}(U\cap X^{o})} with C1πhhQ|πEC2πhC_{1}\pi^{\ast}h\leqslant h^{\prime}_{Q}|_{\pi^{\ast}E}\leqslant C_{2}\pi^{\ast}h and

    (i=1r|πfi|2)chQChQ(\sum^{r}_{i=1}|\pi^{\ast}f_{i}|^{2})^{c}h_{Q}\leqslant Ch^{\prime}_{Q}

    for some c,C,C1,C2+c,C,C_{1},C_{2}\in\mathbb{R}_{+}. Here f1,,frf_{1},...,f_{r} are arbitrary defining functions of UXoU\setminus X^{o}.

In this paper, tame bundles are constructed from harmonic bundles. So we should also recall this notion. Let (H,D)(H,D) be a holomorphic vector bundle on XX with a flat connection DD. Let hHh_{H} be an arbitrary Hermitian metric on HH.

Decompose D=D1,0+D0,1D=D^{1,0}+D^{0,1} into operators of type (1,0)(1,0) and (0,1)(0,1) respectively. Let δH\delta^{\prime}_{H} and δH′′\delta^{\prime\prime}_{H} be the unique operators of type (1,0)(1,0) and (0,1)(0,1) such that the connections D1,0+δHD^{1,0}+\delta^{\prime}_{H} and D0,1+δH′′D^{0,1}+\delta^{\prime\prime}_{H} preserve hHh_{H}. Denote θ=12(D1,0δH)\theta=\frac{1}{2}(D^{1,0}-\delta^{\prime}_{H}), DHc=δH′′δHD^{c}_{H}=\delta^{\prime\prime}_{H}-\delta^{\prime}_{H} and ΘH(D)=DDHc+DHcD\Theta_{H}(D)=DD^{c}_{H}+D^{c}_{H}D.

Remark 2.1.

ΘH(D)\Theta_{H}(D) is called the pseudo-curvature associated with hHh_{H}. In this paper, we will denote the curvature associated with hHh_{H} by ΘH,hH\Theta_{H,h_{H}} to distinguish these two notions.

Definition 2.3.

(H,θ,hH)(H,\theta,h_{H}) is called a harmonic bundle if ΘH(D)=0\Theta_{H}(D)=0. In this case, hHh_{H} is called a harmonic metric.

This notion plays an important role in non-abelian Hodge theory. More precisely, it helps to establish the correspondence between local systems and semistable Higgs bundles with vanishing Chern classes.

Theorem 2.1 ((c.f. [Sim92])).

There is an equivalence of categories between the category of semisimple flat bundles on XX and the category of polystable Higgs bundles with c1(E)=0c_{1}(E)=0 and c2(E)=0c_{2}(E)=0, both being equivalent to the category of harmonic bundles

Now we are ready to reformulate two canonical types of tame bundles. This part should be well-known to experts, and we recommend [ShZ21a] for an explicit exposition.

Proposition 2.2.

Let (H,θ,hH)(H,\theta,h_{H}) be a tame harmonic bundle over XoX^{o}. If EE is a holomorphic subbundle with vanishing second fundamental form and θ¯(E)=0\bar{\theta}(E)=0, then (E,hH|E)(E,h_{H}|_{E}) is a tame hermitian vector bundle with Nakano semi-positive curvature.

In particular, we have

Proposition 2.3.

A real variation of polarized Hodge structure (𝕍,,,S)(\mathbb{V},\nabla,\mathcal{F}^{\cdot},S) on XoX^{o} is a tame harmonic bundle.

3 The harmonic theory

In this section, we develop the harmonic theory concerning the singular Hermitian metrics in order to prove Theorem 1.3. It is mainly inspired by [Tak95]. We prefer to first consider the case that E=𝒪XoE=\mathcal{O}_{X^{o}}, in order to make it easy to understand. However, the whole things should be valid after tensoring with a non-trivial E(h)E(h) as is indicated in [Tak95].

3.1 Background

Let (X,ω)(X,\omega) be a Kähler manifold of dimension nn, and let LL be a holomorphic line bundle on XX endowed with a smooth Hermitian metric φ\varphi. The pointwise inner product ,φ,ω\langle\cdot,\cdot\rangle_{\varphi,\omega} on Ap,q(X,L)A^{p,q}(X,L) is defined by the equation:

α,βφ,ωdVω:=αβ¯eφ\langle\alpha,\beta\rangle_{\varphi,\omega}dV_{\omega}:=\alpha\wedge\ast\bar{\beta}e^{-\varphi}

for α,βAp,q(X,L)\alpha,\beta\in A^{p,q}(X,L). The L2L^{2}-inner product is defined by

(α,β)φ,ω:=Xα,βφ,ω𝑑Vω(\alpha,\beta)_{\varphi,\omega}:=\int_{X}\langle\alpha,\beta\rangle_{\varphi,\omega}dV_{\omega}

for α,βAp,q(X,L)\alpha,\beta\in A^{p,q}(X,L), and the norm φ,ω\|\cdot\|_{\varphi,\omega} is induced by (,)φ,ω(\cdot,\cdot)_{\varphi,\omega}. The standard operators such as ¯\bar{\partial}, φ\partial_{\varphi}, ¯φ:=φ\bar{\partial}^{\ast}_{\varphi}:=-\ast\partial_{\varphi}\ast as well as LL, Λ\Lambda, etc., in Kähler geometry are defined with respect to (,)φ,ω(\cdot,\cdot)_{\varphi,\omega} on XX. In particular, for a smooth (s,t)(s,t)-form γ\gamma, let e(γ)e(\gamma) be the morphism

γ:Ap,q(X,L)Ap+s,q+t(X,L).\gamma\wedge\cdot:A^{p,q}(X,L)\rightarrow A^{p+s,q+t}(X,L).

We then define the operator on Ap,q(X,L)A^{p,q}(X,L) by e(γ):=(1)(p+q)(s+t+1)e(γ¯)e(\gamma)^{\ast}:=(-1)^{(p+q)(s+t+1)}\ast e(\bar{\gamma})\ast. Obviously, e(γ)e(\gamma)^{\ast} is the adjoint operator of e(γ)e(\gamma) with respect to any metric on LL, with or without the compactness or completeness assumptions on the base manifold.

Next we recall the harmonic theory in a local setting [Tak95]. Let VV be a bounded domain with smooth boundary V\partial V on XX. Assume that there is a smooth plurisubharmonic exhaustion function rr of VV, which is defined on a bigger neighbourhood UU with supV(|r|+|dr|)<\sup_{V}(|r|+|dr|)<\infty. In particular, V={r<0}V=\{r<0\} and dr0dr\neq 0 on V\partial V. The volume form dSdS of the real hypersurface V\partial V is defined by dS:=(dr)/|dr|ωdS:=\ast(dr)/|dr|_{\omega}. Setting τ:=dS/|dr|ω\tau:=dS/|dr|_{\omega} we define the inner product on V\partial V by

[α,β]φ,ω:=Vα,βφ,ωτ[\alpha,\beta]_{\varphi,\omega}:=\int_{\partial V}\langle\alpha,\beta\rangle_{\varphi,\omega}\tau

for α,βAp,q(V¯,L)φ,ω\alpha,\beta\in A^{p,q}(\bar{V},L)_{\varphi,\omega}. Then by Stokes’ theorem we have the following:

(¯α,β)φ,ω=(α,¯φβ)φ,ω+[α,e(¯r)β]φ,ω,(φα,β)φ,ω=(α,φβ)φ,ω+[α,e(r)β]φ,ω.\begin{split}(\bar{\partial}\alpha,\beta)_{\varphi,\omega}&=(\alpha,\bar{\partial}^{\ast}_{\varphi}\beta)_{\varphi,\omega}+[\alpha,e(\bar{\partial}r)^{\ast}\beta]_{\varphi,\omega},\\ (\partial_{\varphi}\alpha,\beta)_{\varphi,\omega}&=(\alpha,\partial^{\ast}_{\varphi}\beta)_{\varphi,\omega}+[\alpha,e(\partial r)^{\ast}\beta]_{\varphi,\omega}.\end{split} (3.1)

The space of harmonic forms on VV is then defined as

n,q(V,L,r):={αAn,q(V,L)φ,ω;¯α=¯φα=e(¯r)α=0}.\mathcal{H}^{n,q}(V,L,r):=\{\alpha\in A^{n,q}(V,L)_{\varphi,\omega};\bar{\partial}\alpha=\bar{\partial}^{\ast}_{\varphi}\alpha=e(\bar{\partial}r)^{\ast}\alpha=0\}.

Start from this space, K. Takegoshi generalised Kollar’s injectivity theorem to the Kähler setting. However, the metrics in our paper are not always smooth. Therefore we will further develop this theory next section so that it is applicable in our case.

In the end, we collect several formulas from [Tak95] for the later reference. The first one is the Calabi–Nakano–Vesentini formula:

φ=¯φ+[iΘL,φ,Λ] for φ:=¯¯φ+¯φ¯and¯φ:=φφ+φφ.\Box_{\varphi}=\overline{\Box}_{\varphi}+[i\Theta_{L,\varphi},\Lambda]\textrm{ for }\Box_{\varphi}:=\bar{\partial}\bar{\partial}^{\ast}_{\varphi}+\bar{\partial}^{\ast}_{\varphi}\bar{\partial}\quad\textrm{and}\quad\overline{\Box}_{\varphi}:=\partial_{\varphi}\partial^{\ast}_{\varphi}+\partial^{\ast}_{\varphi}\partial_{\varphi}. (3.2)

Let ψ\psi be a real-valued smooth function on XX. Replacing the metric by φ+ψ\varphi+\psi, we obtain the following variant:

φ+ψ=¯φ+ψ+[iΘL,φ+i¯ψ,Λ] forφ+ψ:=¯¯φ+ψ+¯φ+ψ¯and¯φ+ψ:=φ+ψφ+ψ+φ+ψφ+ψ.\begin{split}\Box_{\varphi+\psi}&=\overline{\Box}_{\varphi+\psi}+[i\Theta_{L,\varphi}+i\partial\bar{\partial}\psi,\Lambda]\textrm{ for}\\ \Box_{\varphi+\psi}&:=\bar{\partial}\bar{\partial}^{\ast}_{\varphi+\psi}+\bar{\partial}^{\ast}_{\varphi+\psi}\bar{\partial}\quad\textrm{and}\quad\overline{\Box}_{\varphi+\psi}:=\partial_{\varphi+\psi}\partial^{\ast}_{\varphi+\psi}+\partial^{\ast}_{\varphi+\psi}\partial_{\varphi+\psi}.\end{split} (3.3)

Donnelly and Xavier’s formula [DoX84] can be formulated as follows:

[¯,e(¯ψ)]+[φ,e(ψ)]=[ie(¯ψ),Λ].[\bar{\partial},e(\bar{\partial}\psi)^{\ast}]+[\partial^{\ast}_{\varphi},e(\partial\psi)]=[ie(\partial\bar{\partial}\psi),\Lambda]. (3.4)

3.2 The harmonic forms concerning singular metrics

Assume that (L,φ)(L,\varphi) is a pseudo-effective line bundle such that φ\varphi has analytic singularities [Dem12]. In particular, φ\varphi is smooth on VZV\setminus Z for a closed subvariety ZZ. Note that ZZ is at least of real codimension 22. So formula (3.1) is still valid on VZV\setminus Z. The formulas (3.2)-(3.4) are established pointwise and thus make sense on arbitrary Kähler manifold such as XX, VV and VZV\setminus Z.

We construct a complete Kähler metric ωl\omega_{l} on VZV\setminus Z as follows: note that VV is exhausted by a smooth plurisubharmonic function rr, then by [Dem82] it is complete. Moreover, there exists a complete Kähler metric ω~\tilde{\omega} on VZV\setminus Z. Let ωl:=ω+1lω~\omega_{l}:=\omega+\frac{1}{l}\tilde{\omega}. Then ωl\omega_{l} is also complete on VZV\setminus Z for all ll. Moreover, ωl2ωl1ω\omega_{l_{2}}\geqslant\omega_{l_{1}}\geqslant\omega when l2l1l_{2}\leqslant l_{1}. Let l=¯¯φ+¯φ¯\Box_{l}=\bar{\partial}\bar{\partial}^{\ast}_{\varphi}+\bar{\partial}^{\ast}_{\varphi}\bar{\partial} be the Laplacian operator associated to φ,ωl\varphi,\omega_{l}. The harmonic form with respect to φ\varphi is defined as

Definition 3.1.

Let α\alpha be a smooth LL-valued (n,q)(n,q)-form on VV with bounded L2L^{2}-norm with respect to ω,φ\omega,\varphi. Assume that for every l1l\gg 1, there exists αl[α|VZ]\alpha_{l}\in[\alpha|_{V\setminus Z}] such that

  1. (1)

    ¯αl=0\bar{\partial}\alpha_{l}=0, ¯φαl=0\bar{\partial}^{\ast}_{\varphi}\alpha_{l}=0 and e(¯r)αl=0e(\bar{\partial}r)^{\ast}\alpha_{l}=0 on VZV\setminus Z;

  2. (2)

    αlα|VZ\alpha_{l}\rightarrow\alpha|_{V\setminus Z} in the sense of L2L^{2}-norm.

Then we call α\alpha a φ\Box_{\varphi}-harmonic form. The space of all the φ\Box_{\varphi}-harmonic forms is denoted by

n,q(V,L(φ),r).\mathcal{H}^{n,q}(V,L\otimes\mathscr{I}(\varphi),r).

Here αl[α|VZ]\alpha_{l}\in[\alpha|_{V\setminus Z}] means that there exits an LL-valued (n,q1)(n,q-1)-form βl\beta_{l} on VZV\setminus Z, such that α|VZ=αl+¯βl\alpha|_{V\setminus Z}=\alpha_{l}+\bar{\partial}\beta_{l}.

We then generalise several propositions in [Tak95] here.

Proposition 3.1.

We have the following conclusions:

  1. 1.

    Assume αAn,q(VZ,L)φ,ωl\alpha\in A^{n,q}(V\setminus Z,L)_{\varphi,\omega_{l}} satisfies e(¯r)α=0e(\bar{\partial}r)^{\ast}\alpha=0. Then α\alpha satisfies ¯α=¯φα=0\bar{\partial}\alpha=\bar{\partial}^{\ast}_{\varphi}\alpha=0 if and only if ¯α=0\bar{\partial}\ast\alpha=0 and (ie(ΘL,φ+¯r)Λα,α)φ,ωl=0(ie(\Theta_{L,\varphi}+\partial\bar{\partial}r)\Lambda\alpha,\alpha)_{\varphi,\omega_{l}}=0.

  2. 2.

    n,q(V,L(φ),r)\mathcal{H}^{n,q}(V,L\otimes\mathscr{I}(\varphi),r) is independent of the choice of exhaustion function rr.

Proof.

The proof uses the same argument as Theorem 4.3 in [Tak95] with minor adjustment. So we only provide the necessary details. Let Vc:={r<c}V_{c}:=\{r<c\} for every c(,0)c\in(-\infty,0).

(i) Take any regular value cc of rr. If ¯α=¯φα=0\bar{\partial}\alpha=\bar{\partial}^{\ast}_{\varphi}\alpha=0, by formulas (3.1) and (3.2) we obtain

φαφ,ωl2+(iΘL,φΛα,α)φ,ωl+[φα,e(r)α]φ,ωl=0\|\partial^{\ast}_{\varphi}\alpha\|^{2}_{\varphi,\omega_{l}}+(i\Theta_{L,\varphi}\Lambda\alpha,\alpha)_{\varphi,\omega_{l}}+[\partial^{\ast}_{\varphi}\alpha,e(\partial r)^{\ast}\alpha]_{\varphi,\omega_{l}}=0

on VcZV_{c}\setminus Z. Hence we obtain by (3.4)

φαφ,ωl2+(iΘL,φΛα,α)φ,ωl+[ie(¯r)Λα,α]φ,ωl=0\|\partial^{\ast}_{\varphi}\alpha\|^{2}_{\varphi,\omega_{l}}+(i\Theta_{L,\varphi}\Lambda\alpha,\alpha)_{\varphi,\omega_{l}}+[ie(\partial\bar{\partial}r)\Lambda\alpha,\alpha]_{\varphi,\omega_{l}}=0

on VcZV_{c}\setminus Z, which implies φα=¯α=0\partial^{\ast}_{\varphi}\alpha=\ast\bar{\partial}\ast\alpha=0 and (iΘL,φΛα,α)φ,ωl=[ie(¯r)Λα,α]φ,ωl=0(i\Theta_{L,\varphi}\Lambda\alpha,\alpha)_{\varphi,\omega_{l}}=[ie(\partial\bar{\partial}r)\Lambda\alpha,\alpha]_{\varphi,\omega_{l}}=0 by the plurisubharmonicity of rr. Apply (3.4) again we obtain

(ie(¯r)Λα,α)φ,ωl=(ie(r)φα,α)φ,ωl=0.(ie(\partial\bar{\partial}r)\Lambda\alpha,\alpha)_{\varphi,\omega_{l}}=(ie(\partial r)\partial^{\ast}_{\varphi}\alpha,\alpha)_{\varphi,\omega_{l}}=0.

The necessity is then proved as cc varies.

If ¯α=0\bar{\partial}\ast\alpha=0 and (ie(ΘL,φ+¯r)Λα,α)φ,ωl=0(ie(\Theta_{L,\varphi}+\partial\bar{\partial}r)\Lambda\alpha,\alpha)_{\varphi,\omega_{l}}=0, by the plurisubharmonicity of rr

(iΘL,φΛα,α)φ,ωl=0and(ie(¯r)Λα,α)φ,ωl=0(i\Theta_{L,\varphi}\Lambda\alpha,\alpha)_{\varphi,\omega_{l}}=0\quad\textrm{and}\quad(ie(\partial\bar{\partial}r)\Lambda\alpha,\alpha)_{\varphi,\omega_{l}}=0

on VZV\setminus Z. By (3.2) and (3.4) we obtain:

(lα,α)φ,ωl=0and(e(¯r)¯α,α)φ,ωl=0(\Box_{l}\alpha,\alpha)_{\varphi,\omega_{l}}=0\quad\textrm{and}\quad(e(\bar{\partial}r)^{\ast}\bar{\partial}\alpha,\alpha)_{\varphi,\omega_{l}}=0

on VcZV_{c}\setminus Z. Hence by (3.1) we obtain:

0=(lα,α)φ,ωl=¯αφ,ωl2+¯φαφ,ωl20=(\Box_{l}\alpha,\alpha)_{\varphi,\omega_{l}}=\|\bar{\partial}\alpha\|^{2}_{\varphi,\omega_{l}}+\|\bar{\partial}^{\ast}_{\varphi}\alpha\|^{2}_{\varphi,\omega_{l}}

on VcZV_{c}\setminus Z, which implies the sufficiency as cc varies.

(ii) Let τ\tau be an arbitrary smooth plurisubharmonic function defined on a bigger neighbourhood UU containing VV. Given

αn,q(V,L(φ),r),\alpha\in\mathcal{H}^{n,q}(V,L\otimes\mathscr{I}(\varphi),r),

there exits {αl}\{\alpha_{l}\} on VZV\setminus Z with ¯αl=¯φαl=e(¯r)αl=0\bar{\partial}\alpha_{l}=\bar{\partial}^{\ast}_{\varphi}\alpha_{l}=e(\bar{\partial}r)^{\ast}\alpha_{l}=0, and is convergent to α|VZ\alpha|_{V\setminus Z} by definition. Formula (3.4) implies that ¯(e(¯τ)αl)+e(τ)φαl=ie(¯τ)Λαl\bar{\partial}(e(\bar{\partial}\tau)^{\ast}\alpha_{l})+e(\partial\tau)\partial^{\ast}_{\varphi}\alpha_{l}=ie(\partial\bar{\partial}\tau)\Lambda\alpha_{l} on VZV\setminus Z. Therefore

(ie(¯τ)Λαl,αl)φτ,ωl=(¯(e(¯τ)αl),αl)φτ,ωl+(e(τ)φαl,αl)φτ,ωl=(e(¯τ)αl,¯φταl)φτ,ωl+(e(τ)φαl,αl)φτ,ωl=e(¯τ)αlφτ,ωl2+(e(τ)φαl,αl)φτ,ωl\begin{split}&(ie(\partial\bar{\partial}\tau)\Lambda\alpha_{l},\alpha_{l})_{\varphi-\tau,\omega_{l}}\\ =&(\bar{\partial}(e(\bar{\partial}\tau)^{\ast}\alpha_{l}),\alpha_{l})_{\varphi-\tau,\omega_{l}}+(e(\partial\tau)\partial^{\ast}_{\varphi}\alpha_{l},\alpha_{l})_{\varphi-\tau,\omega_{l}}\\ =&(e(\bar{\partial}\tau)^{\ast}\alpha_{l},\bar{\partial}^{\ast}_{\varphi-\tau}\alpha_{l})_{\varphi-\tau,\omega_{l}}+(e(\partial\tau)\partial^{\ast}_{\varphi}\alpha_{l},\alpha_{l})_{\varphi-\tau,\omega_{l}}\\ =&-\|e(\bar{\partial}\tau)^{\ast}\alpha_{l}\|^{2}_{\varphi-\tau,\omega_{l}}+(e(\partial\tau)\partial^{\ast}_{\varphi}\alpha_{l},\alpha_{l})_{\varphi-\tau,\omega_{l}}\end{split}

on VcZV_{c}\setminus Z for any non-critical value cc of rr. Note φαl=0\partial^{\ast}_{\varphi}\alpha_{l}=0 by (i). We then obtain that

(ie(¯τ)Λαl,αl)φτ,ωl=e(¯τ)αlφτ,ωl2 on VcZ.(ie(\partial\bar{\partial}\tau)\Lambda\alpha_{l},\alpha_{l})_{\varphi-\tau,\omega_{l}}=-\|e(\bar{\partial}\tau)^{\ast}\alpha_{l}\|^{2}_{\varphi-\tau,\omega_{l}}\textrm{ on }V_{c}\setminus Z.

Notice that τ\tau is plurisubharmonic, as cc varies we actually have

(ie(¯τ)Λαl,αl)φτ,ωl=e(¯τ)αlφτ,ωl2=0 on VZ.(ie(\partial\bar{\partial}\tau)\Lambda\alpha_{l},\alpha_{l})_{\varphi-\tau,\omega_{l}}=\|e(\bar{\partial}\tau)^{\ast}\alpha_{l}\|^{2}_{\varphi-\tau,\omega_{l}}=0\textrm{ on }V\setminus Z.

Obviously it is equivalent to say that

(ie(¯τ)Λαl,αl)φ,ωl=e(¯τ)αlφ,ωl2=0 on VZ.(ie(\partial\bar{\partial}\tau)\Lambda\alpha_{l},\alpha_{l})_{\varphi,\omega_{l}}=\|e(\bar{\partial}\tau)^{\ast}\alpha_{l}\|^{2}_{\varphi,\omega_{l}}=0\textrm{ on }V\setminus Z.

Combine with (i), we eventually obtain that

n,q(V,L(φ),r)=n,q(V,L(φ),r+τ)\mathcal{H}^{n,q}(V,L\otimes\mathscr{I}(\varphi),r)=\mathcal{H}^{n,q}(V,L\otimes\mathscr{I}(\varphi),r+\tau)

for any smooth plurisubharmonic τ\tau, hence the desired conclusion. ∎

3.3 The Hodge-type isomorphism

In this section, we return to the relative setting. Let f:XYf:X\rightarrow Y be a proper, locally Kähler morphism from a complex manifold XX to a reduced, irreducible analytic space YY. Every connected component of XX is mapped surjectively to YY. Let ll be the dimension of a general fibre FF of ff.

Suppose that (L,φ)(L,\varphi) is a pseudo-effective line bundle on XX with analytic singularities. Let ZZ be the closed subvariety such that φ\varphi is smooth on XZX\setminus Z. Let {U}\{U\} be a Stein covering of YY, then f1(U)f^{-1}(U) is Kähler. In particular, we could construct complete Kähler metrics ωl\omega_{l} on f1(U)Zf^{-1}(U)\setminus Z as is shown before.

Let rUr_{U} be a smooth strictly plurisubharmonic exhaustion function on UU. In particular,

supU(|rU|+|drU|)<\sup_{U}(|r_{U}|+|dr_{U}|)<\infty

after shrinking UU if necessary. Let

n,q(f1(U),L(φ),frU)\mathcal{H}^{n,q}(f^{-1}(U),L\otimes\mathscr{I}(\varphi),f^{\ast}r_{U})

be the harmonic space in Definition 3.1. We have

Proposition 3.2.
  1. 1.

    n,q(f1(U),L(φ),frU)Hq(f1(U),KXL(φ))\mathcal{H}^{n,q}(f^{-1}(U),L\otimes\mathscr{I}(\varphi),f^{\ast}r_{U})\simeq H^{q}(f^{-1}(U),K_{X}\otimes L\otimes\mathscr{I}(\varphi)).

  2. 2.

    If VV is a Stein open subset of UU provided with a smooth strictly plurisubharmonic exhaustion function rVr_{V}, then the restriction map

    n,q(f1(U),L(φ),frU)n,q(f1(V),(φ),frV)\mathcal{H}^{n,q}(f^{-1}(U),L\otimes\mathscr{I}(\varphi),f^{\ast}r_{U})\rightarrow\mathcal{H}^{n,q}(f^{-1}(V),\mathscr{I}(\varphi),f^{\ast}r_{V})

    is well-defined, and further the following diagram is commutative:

    n,q(f1(U),L(φ),frU)\textstyle{\mathcal{H}^{n,q}(f^{-1}(U),L\otimes\mathscr{I}(\varphi),f^{\ast}r_{U})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}iVU\scriptstyle{i^{U}_{V}}SUq\scriptstyle{S^{q}_{U}}H0(f1(U),ΩXnqL(φ))\textstyle{H^{0}(f^{-1}(U),\Omega^{n-q}_{X}\otimes L\otimes\mathscr{I}(\varphi))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}n,q(f1(V),L(φ),frV)\textstyle{\mathcal{H}^{n,q}(f^{-1}(V),L\otimes\mathscr{I}(\varphi),f^{\ast}r_{V})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}SVq\scriptstyle{S^{q}_{V}}H0(f1(V),ΩXnqL(φ)).\textstyle{H^{0}(f^{-1}(V),\Omega^{n-q}_{X}\otimes L\otimes\mathscr{I}(\varphi)).}

    The morphism SUqS^{q}_{U} will be verified during the proof.

Proof.

(i) When φ\varphi is smooth, it is nothing but [Tak95], Theorem 5.2. Whereas our φ\varphi is merely smooth outside ZZ, then the idea is to apply the similar argument on f1(U)Zf^{-1}(U)\setminus Z to obtain the desired conclusion.

More precisely, let L(2)n,q(f1(U),L)φ,ωL^{n,q}_{(2)}(f^{-1}(U),L)_{\varphi,\omega} be the space of LL-valued (n,q)(n,q)-forms on f1(U)f^{-1}(U) which are L2L^{2}-bounded with respect to φ,ω\varphi,\omega. Let [α]Hq(f1(U),KXL(φ))[\alpha]\in H^{q}(f^{-1}(U),K_{X}\otimes L\otimes\mathscr{I}(\varphi)). We use the de Rham–Weil isomorphism

Hq(f1(U),KXL(φ))Ker(¯:L(2)n,q(f1(U),L)φ,ωL(2)n,q+1(f1(U),L)φ,ω)Im(¯:L(2)n,q1(f1(U),L)φ,ωL(2)n,q(f1(U),L)φ,ω)H^{q}(f^{-1}(U),K_{X}\otimes L\otimes\mathscr{I}(\varphi))\simeq\frac{\mathrm{Ker}(\bar{\partial}:L^{n,q}_{(2)}(f^{-1}(U),L)_{\varphi,\omega}\rightarrow L^{n,q+1}_{(2)}(f^{-1}(U),L)_{\varphi,\omega})}{\mathrm{Im}(\bar{\partial}:L^{n,q-1}_{(2)}(f^{-1}(U),L)_{\varphi,\omega}\rightarrow L^{n,q}_{(2)}(f^{-1}(U),L)_{\varphi,\omega})}

to represent it by a ¯\bar{\partial}-closed LL-valued (n,q)(n,q)-form α\alpha with αφ,ω<\|\alpha\|_{\varphi,\omega}<\infty.

Let L(2)n,q(f1(U)Z,L)φ,ωlL^{n,q}_{(2)}(f^{-1}(U)\setminus Z,L)_{\varphi,\omega_{l}} be the space of LL-valued (n,q)(n,q)-forms on f1(U)Zf^{-1}(U)\setminus Z which are L2L^{2}-bounded with respect to φ,ωl\varphi,\omega_{l}. Since φ\varphi is smooth on f1(U)Zf^{-1}(U)\setminus Z, we have the orthogonal decomposition

L(2)n,q(f1(U)Z,L)φ,ωl=Im¯¯φ,ωln,q(L)Im¯φ¯,L^{n,q}_{(2)}(f^{-1}(U)\setminus Z,L)_{\varphi,\omega_{l}}=\overline{\mathrm{Im}\bar{\partial}}\bigoplus\mathcal{H}^{n,q}_{\varphi,\omega_{l}}(L)\bigoplus\overline{\mathrm{Im}\bar{\partial}^{\ast}_{\varphi}},

where

φ,ωln,q(L)={αDom¯Dom¯φ;¯α=0,¯φα=0}\mathcal{H}^{n,q}_{\varphi,\omega_{l}}(L)=\{\alpha\in\textrm{Dom}\bar{\partial}\cap\textrm{Dom}\bar{\partial}^{\ast}_{\varphi};\bar{\partial}\alpha=0,\bar{\partial}^{\ast}_{\varphi}\alpha=0\}

with

Dom¯={αL(2)n,q(f1(U)Z,L)φ,ωl;¯αL(2)n,q+1(f1(U)Z,L)φ,ωl}\textrm{Dom}\bar{\partial}=\{\alpha\in L^{n,q}_{(2)}(f^{-1}(U)\setminus Z,L)_{\varphi,\omega_{l}};\bar{\partial}\alpha\in L^{n,q+1}_{(2)}(f^{-1}(U)\setminus Z,L)_{\varphi,\omega_{l}}\}

and

Dom¯φ={αL(2)n,q(f1(U)Z,L)φ,ωl;¯φαL(2)n,q1(f1(U)Z,L)φ,ωl}.\textrm{Dom}\bar{\partial}^{\ast}_{\varphi}=\{\alpha\in L^{n,q}_{(2)}(f^{-1}(U)\setminus Z,L)_{\varphi,\omega_{l}};\bar{\partial}^{\ast}_{\varphi}\alpha\in L^{n,q-1}_{(2)}(f^{-1}(U)\setminus Z,L)_{\varphi,\omega_{l}}\}.

Moreover, for any uIm¯¯An,q(f1(U)Z,L)u\in\overline{\mathrm{Im}\bar{\partial}}\cap A^{n,q}(f^{-1}(U)\setminus Z,L), there is vAn,q1(f1(U)Z,L)v\in A^{n,q-1}(f^{-1}(U)\setminus Z,L) such that u=¯vu=\bar{\partial}v. It is essentially due to the fact that ff is proper and

Hq(f1(U)Z,KXL)H^{q}(f^{-1}(U)\setminus Z,K_{X}\otimes L)

is a separated topological space. We recommend [Tak95], Proposition 4.6 or [Fuj12], Claim 1 for references. Now we set αl:=\alpha_{l}:= the orthogonal projection of α|f1(U)Z\alpha|_{f^{-1}(U)\setminus Z} onto φ,ωln,q(L)\mathcal{H}^{n,q}_{\varphi,\omega_{l}}(L), which is smooth by the regularization theorem for elliptic operators of second order. Hence

α|f1(U)ZαlIm¯¯An,q(f1(U)Z,L).\alpha|_{f^{-1}(U)\setminus Z}-\alpha_{l}\in\overline{\mathrm{Im}\bar{\partial}}\cap A^{n,q}(f^{-1}(U)\setminus Z,L).

Then there is βlAn,q1(f1(U)Z,L)\beta_{l}\in A^{n,q-1}(f^{-1}(U)\setminus Z,L) such that α|f1(U)Z=αl+¯βl\alpha|_{f^{-1}(U)\setminus Z}=\alpha_{l}+\bar{\partial}\beta_{l}. Moreover, the representative

αl[α|f1(U)Z]\alpha_{l}\in[\alpha|_{f^{-1}(U)\setminus Z}]

minimizes the L2L^{2}-norm defined by φ\varphi and ωl\omega_{l}. Now we claim that e(¯frU)αl=0e(\bar{\partial}f^{\ast}r_{U})^{\ast}\alpha_{l}=0.

In order to prove this claim, we first recall the following formula in [Tak95], Theorem 2.2. Let ψ\psi be a smooth plurisubharmonic function on f1(U)Zf^{-1}(U)\setminus Z such that

supf1(U)Z(|ψ|+|dψ|)<\sup_{f^{-1}(U)\setminus Z}(|\psi|+|d\psi|)<\infty

and let η=eψ\eta=e^{\psi}, then

η(¯+e(¯ψ))βφ,ωl2+η¯φβφ,ωl2=η(φe(ψ))βφ,ωl2+(iη(ΘL,φ+¯ψ)Λβ,β)φ,ωl\begin{split}&\|\sqrt{\eta}(\bar{\partial}+e(\bar{\partial}\psi))\beta\|^{2}_{\varphi,\omega_{l}}+\|\sqrt{\eta}\bar{\partial}^{\ast}_{\varphi}\beta\|^{2}_{\varphi,\omega_{l}}\\ =&\|\sqrt{\eta}(\partial^{\ast}_{\varphi}-e(\partial\psi))^{\ast}\beta\|^{2}_{\varphi,\omega_{l}}+(i\eta(\Theta_{L,\varphi}+\partial\bar{\partial}\psi)\Lambda\beta,\beta)_{\varphi,\omega_{l}}\end{split} (3.5)

for any βDom¯Dom¯φL(2)n,q(f1(U)Z,L)φ,ωl\beta\in\textrm{Dom}\bar{\partial}\cap\textrm{Dom}\bar{\partial}^{\ast}_{\varphi}\subseteq L^{n,q}_{(2)}(f^{-1}(U)\setminus Z,L)_{\varphi,\omega_{l}}. We apply (3.5) with β=αl\beta=\alpha_{l} and ψ=0\psi=0, then

0=φαlφ,ωl2+(iΘL,φΛαl,αl)φ,ωl.0=\|\partial^{\ast}_{\varphi}\alpha_{l}\|^{2}_{\varphi,\omega_{l}}+(i\Theta_{L,\varphi}\Lambda\alpha_{l},\alpha_{l})_{\varphi,\omega_{l}}.

Thus, we obtain that

φαl=0and(iΘL,φΛαl,αl)φ,ωl=0.\partial^{\ast}_{\varphi}\alpha_{l}=0\quad\textrm{and}\quad(i\Theta_{L,\varphi}\Lambda\alpha_{l},\alpha_{l})_{\varphi,\omega_{l}}=0.

Apply (3.5) again with β=αl\beta=\alpha_{l} and ψ=frU\psi=f^{\ast}r_{U}, we can obtain

e(¯frU)αlφfrU,ωl2=e(frU)αlφfrU,ωl2+(ie(¯frU)Λαl,αl)φfrU,ωl.\|e(\bar{\partial}f^{\ast}r_{U})\alpha_{l}\|^{2}_{\varphi-f^{\ast}r_{U},\omega_{l}}=\|e(\partial f^{\ast}r_{U})^{\ast}\alpha_{l}\|^{2}_{\varphi-f^{\ast}r_{U},\omega_{l}}+(ie(\partial\bar{\partial}f^{\ast}r_{U})\Lambda\alpha_{l},\alpha_{l})_{\varphi-f^{\ast}r_{U},\omega_{l}}. (3.6)

On the other hand, it is easy to deduce the following equation:

e(frU)αlφfrU,ωl2=e(¯frU)αlφfrU,ωl2+e(¯frU)αlφfrU,ωl2\|e(\partial f^{\ast}r_{U})^{\ast}\alpha_{l}\|^{2}_{\varphi-f^{\ast}r_{U},\omega_{l}}=\|e(\bar{\partial}f^{\ast}r_{U})\alpha_{l}\|^{2}_{\varphi-f^{\ast}r_{U},\omega_{l}}+\|e(\bar{\partial}f^{\ast}r_{U})^{\ast}\alpha_{l}\|^{2}_{\varphi-f^{\ast}r_{U},\omega_{l}}

on f1(U)Zf^{-1}(U)\setminus Z from the Kähler identity:

e(θ)=i[e(θ¯),Λ]e(\theta)^{\ast}=i[e(\bar{\theta}),\Lambda]

for θA1,0(f1(U)Z)\theta\in A^{1,0}(f^{-1}(U)\setminus Z). Finally we obtain that e(¯frU)αl=0e(\bar{\partial}f^{\ast}r_{U})^{\ast}\alpha_{l}=0 from the plurisubharmonicity of frUf^{\ast}r_{U}. The claim is then proved. It remains to show that {αl}\{\alpha_{l}\} is convergent to α^|f1(U)Z\hat{\alpha}|_{f^{-1}(U)\setminus Z} for some smooth LL-valued (n,q)(n,q)-form α^\hat{\alpha} on f1(U)f^{-1}(U).

Fix l0l_{0}. For any ll0l\geqslant l_{0} we have

αlφ,ωl0αlφ,ωl=αlφ,ωlα|f1(U)Zφ,ωlαφ,ω<.\|\ast\alpha_{l}\|_{\varphi,\omega_{l_{0}}}\leqslant\|\ast\alpha_{l}\|_{\varphi,\omega_{l}}=\|\alpha_{l}\|_{\varphi,\omega_{l}}\leqslant\|\alpha|_{f^{-1}(U)\setminus Z}\|_{\varphi,\omega_{l}}\leqslant\|\alpha\|_{\varphi,\omega}<\infty.

Here we use the fact that αl\alpha_{l} minimizes the L2L^{2}-norm φ,ωl\|\cdot\|_{\varphi,\omega_{l}}. It means that αl\ast\alpha_{l} has uniformly bounded L2L^{2}-norm with respect to φ,ωl0\|\cdot\|_{\varphi,\omega_{l_{0}}} on f1(U)Zf^{-1}(U)\setminus Z. On the other hand, since

0=φαl=¯αl,0=\partial^{\ast}_{\varphi}\alpha_{l}=\ast\bar{\partial}\ast\alpha_{l},

αl\ast\alpha_{l} is a holomorphic LL-valued (nq,0)(n-q,0)-form on f1(U)Zf^{-1}(U)\setminus Z. Therefore, it extends as θl\theta_{l} to the whole space by classic L2L^{2}-extension theorem [Ohs02]. Due to the fact that θlφ,ωl0\|\theta_{l}\|_{\varphi,\omega_{l_{0}}} is uniformly bounded, {θl}\{\theta_{l}\} converges to some θ\theta in the sense of L2L^{2}-norm φ,ωl0\|\cdot\|_{\varphi,\omega_{l_{0}}}. Since θl\theta_{l} is holomorphic, so is θ\theta. It means that θl\theta_{l} is actually convergent to θ\theta with respect to any L2L^{2}-norm. Now, since θφ,ωl0αφ,ω\|\theta\|_{\varphi,\omega_{l_{0}}}\leqslant\|\alpha\|_{\varphi,\omega}, we actually have θφ,ωαφ,ω\|\theta\|_{\varphi,\omega}\leqslant\|\alpha\|_{\varphi,\omega} as l0l_{0} tends to infinity. In summary, we have defined a morphism as follows so far:

SUq:Hq(f1(U),KXL(φ))H0(f1(U),ΩXnqL(φ))[α]θ.\begin{split}S^{q}_{U}:H^{q}(f^{-1}(U),K_{X}\otimes L\otimes\mathscr{I}(\varphi))&\rightarrow H^{0}(f^{-1}(U),\Omega^{n-q}_{X}\otimes L\otimes\mathscr{I}(\varphi))\\ [\alpha]&\mapsto\theta.\end{split}

Now let cnq=(i)(nq)2c_{n-q}=(i)^{(n-q)^{2}}, and let γl=cnqωlqq!θl\gamma_{l}=c_{n-q}\frac{\omega^{q}_{l}}{q!}\wedge\theta_{l}. Obviously we have

cnqωqq!θ=limlγl=limlcnqωlqq!αl=limlαl on f1(U)Z.c_{n-q}\frac{\omega^{q}}{q!}\wedge\theta=\lim_{l\rightarrow\infty}\gamma_{l}=\lim_{l\rightarrow\infty}c_{n-q}\frac{\omega^{q}_{l}}{q!}\wedge\ast\alpha_{l}=\lim_{l\rightarrow\infty}\alpha_{l}\textrm{ on }f^{-1}(U)\setminus Z. (3.7)

Let α^:=cnqωqq!θ\hat{\alpha}:=c_{n-q}\frac{\omega^{q}}{q!}\wedge\theta, then it is a smooth LL-valued (n,q)(n,q)-form on f1(U)f^{-1}(U) with bounded L2L^{2}-norm φ,ω\|\cdot\|_{\varphi,\omega}. Now

α^n,q(f1(U),L(φ),frU)\hat{\alpha}\in\mathcal{H}^{n,q}(f^{-1}(U),L\otimes\mathscr{I}(\varphi),f^{\ast}r_{U})

by definition. We denote this morphism by i([α])=α^i([\alpha])=\hat{\alpha}.

Conversely, for an αn,q(f1(U),L(φ),frU)\alpha\in\mathcal{H}^{n,q}(f^{-1}(U),L\otimes\mathscr{I}(\varphi),f^{\ast}r_{U}), by definition there exists an

αl[α|f1(U)Z]\alpha_{l}\in[\alpha|_{f^{-1}(U)\setminus Z}]

with αlφ,ωln,q(L)\alpha_{l}\in\mathcal{H}^{n,q}_{\varphi,\omega_{l}}(L) for every ll and limαl=α|f1(U)Z\lim\alpha_{l}=\alpha|_{f^{-1}(U)\setminus Z}. In particular, ¯αl=0\bar{\partial}\alpha_{l}=0. So

¯(α|f1(U)Z)=0.\bar{\partial}(\alpha|_{f^{-1}(U)\setminus Z})=0.

Since α\alpha is smooth, we actually have ¯α=0\bar{\partial}\alpha=0 on f1(U)f^{-1}(U). It then defines a cohomology class

[α]Hq(f1(U),KXL(φ)).[\alpha]\in H^{q}(f^{-1}(U),K_{X}\otimes L\otimes\mathscr{I}(\varphi)).

We denote this morphism by j(α)=[α]j(\alpha)=[\alpha]. It is easy to verify that ij=idi\circ j=\textrm{id} and ji=idj\circ i=\textrm{id}. The proof is finished.

(ii) Let αn,q(f1(U),L(φ),frU)\alpha\in\mathcal{H}^{n,q}(f^{-1}(U),L\otimes\mathscr{I}(\varphi),f^{\ast}r_{U}), and let {αl}\{\alpha_{l}\} be the sequence on f1(U)Zf^{-1}(U)\setminus Z that is convergent to α|f1(U)Z\alpha|_{f^{-1}(U)\setminus Z}. Obviously

αl|f1(V)Zφ,ωln,q(L)L(2)n,q(f1(V)Z,L)φ,ωl.\alpha_{l}|_{f^{-1}(V)\setminus Z}\in\mathcal{H}^{n,q}_{\varphi,\omega_{l}}(L)\subseteq L^{n,q}_{(2)}(f^{-1}(V)\setminus Z,L)_{\varphi,\omega_{l}}.

In order to show the restriction map is well-defined, we only need to prove that

e(¯frV)αl|f1(V)Z=0.e(\bar{\partial}f^{\ast}r_{V})^{\ast}\alpha_{l}|_{f^{-1}(V)\setminus Z}=0.

However, it is nothing but repeat to argument in (i). Technically, apply (3.5) on f1(V)Zf^{-1}(V)\setminus Z we obtain that

φαl=0and(iΘL,φΛαl,αl)φ,ωl=0.\partial^{\ast}_{\varphi}\alpha_{l}=0\quad\textrm{and}\quad(i\Theta_{L,\varphi}\Lambda\alpha_{l},\alpha_{l})_{\varphi,\omega_{l}}=0.

Apply (3.5) again with β=αl\beta=\alpha_{l} and ψ=frV\psi=f^{\ast}r_{V}, we can obtain

e(¯frV)αlφfrV,ωl2=e(frV)αlφfrV,ωl2+(iηe(¯frV)Λαl,αl)φ,ωl.\begin{split}\|e(\bar{\partial}f^{\ast}r_{V})\alpha_{l}\|^{2}_{\varphi-f^{\ast}r_{V},\omega_{l}}=\|e(\partial f^{\ast}r_{V})^{\ast}\alpha_{l}\|^{2}_{\varphi-f^{\ast}r_{V},\omega_{l}}+(i\eta e(\partial\bar{\partial}f^{\ast}r_{V})\Lambda\alpha_{l},\alpha_{l})_{\varphi,\omega_{l}}.\end{split} (3.8)

Combine with the following equation:

e(frV)αlφfrV,ωl2=e(¯frV)αlφfrV,ωl2+e(¯frV)αlφfrV,ωl2,\|e(\partial f^{\ast}r_{V})^{\ast}\alpha_{l}\|^{2}_{\varphi-f^{\ast}r_{V},\omega_{l}}=\|e(\bar{\partial}f^{\ast}r_{V})\alpha_{l}\|^{2}_{\varphi-f^{\ast}r_{V},\omega_{l}}+\|e(\bar{\partial}f^{\ast}r_{V})^{\ast}\alpha_{l}\|^{2}_{\varphi-f^{\ast}r_{V},\omega_{l}},

finally we obtain that e(¯frV)αl=0e(\bar{\partial}f^{\ast}r_{V})^{\ast}\alpha_{l}=0 from the plurisubharmonicity of frVf^{\ast}r_{V}. Then everything is intuitive due to the discussions before. ∎

The data

{n,q(f1(U),L(φ),frU),iVU}\{\mathcal{H}^{n,q}(f^{-1}(U),L\otimes\mathscr{I}(\varphi),f^{\ast}r_{U}),i^{U}_{V}\}

with the restriction morphisms

iVU:n,q(f1(U),L(φ),frU)n,q(f1(V),L(φ),frV),i^{U}_{V}:\mathcal{H}^{n,q}(f^{-1}(U),L\otimes\mathscr{I}(\varphi),f^{\ast}r_{U})\rightarrow\mathcal{H}^{n,q}(f^{-1}(V),L\otimes\mathscr{I}(\varphi),f^{\ast}r_{V}),

(V,rV)(U,rU)(V,r_{V})\subset(U,r_{U}), yields a presheaf [Har77] on YY by Proposition 3.2, (ii). We denote the associated sheaf by fn,q(L(φ))f_{\ast}\mathcal{H}^{n,q}(L\otimes\mathscr{I}(\varphi)). Since

Rqf(KXL(φ))R^{q}f_{\ast}(K_{X}\otimes L\otimes\mathscr{I}(\varphi))

is defined as the sheaf associated with the presheaf

UHq(f1(U),KXL(φ)),U\rightarrow H^{q}(f^{-1}(U),K_{X}\otimes L\otimes\mathscr{I}(\varphi)),

the sheaf fn,q(L(φ))f_{\ast}\mathcal{H}^{n,q}(L\otimes\mathscr{I}(\varphi)) is isomorphic to Rqf(KXL(φ))R^{q}f_{\ast}(K_{X}\otimes L\otimes\mathscr{I}(\varphi)) by Proposition 3.2, (i).

Remark 3.1.

The whole harmonic theory could even be established for a pseudo-effective line bundle (L,φ)(L,\varphi) with following property: there exists integers k0k_{0}, mm and sections s1,,smLk0s_{1},...,s_{m}\in L^{k_{0}} such that

supX(i=1m|si|2eφ)<.\sup_{X}(\sum^{m}_{i=1}|s_{i}|^{2}e^{-\varphi})<\infty.

Readers may refer to [Wu20] for more details.

Remark 3.2.

Everything in this section works smoothly after tensoring with E(h)E(h) provided that (E,h)(E,h) is a tame Nakano semi-positive vector bundle on XoX^{o}. Here XoX^{o} is a Zariski open subset of XX. We list the major adjustments along with brief explanations here.

  1. (1)

    In Definition 3.1, VZV\setminus Z is replaced by (VZ)Xo(V\setminus Z)\cap X^{o}, which is still complete for the same reason. The L2L^{2}-norm is defined by ω,φ,h\omega,\varphi,h;

  2. (2)

    Since XoX^{o} is Zariski open, formulas (3.1)-(3.5) still hold except that the curvature term is exchanged as iΘLE,φ,hi\Theta_{L\otimes E,\varphi,h}. It is semi-positive in the sense of Nakano hence the argument in Propositions 3.1 and 3.2 extends with no difficulties.

  3. (3)

    fn,q(L(φ))f_{\ast}\mathcal{H}^{n,q}(L\otimes\mathscr{I}(\varphi)) is replaced by fn,q(LE(heφ))f_{\ast}\mathcal{H}^{n,q}(L\otimes E(h\cdot e^{-\varphi})), and Rqf(KXL(φ))R^{q}f_{\ast}(K_{X}\otimes L\otimes\mathscr{I}(\varphi)) is replaced by Rqf(KXLE(heφ))R^{q}f_{\ast}(K_{X}\otimes L\otimes E(h\cdot e^{-\varphi})).

Now we will directly use the following isomorphism:

Rqf(KXLE(heφ))fn,q(LE(heφ)).R^{q}f_{\ast}(K_{X}\otimes L\otimes E(h\cdot e^{-\varphi}))\simeq f_{\ast}\mathcal{H}^{n,q}(L\otimes E(h\cdot e^{-\varphi})). (3.9)

This should not lead to any confusion.

In the end, remember that for a relative asymptotic multiplier ideal sheaf (f,L)\mathscr{I}(f,\|L\|), we can always associate a collection of singular metrics {f1(U),φU}\{f^{-1}(U),\varphi_{U}\} with algebraic singularities [Dem12] on LL, such that

iΘL,φU0and(φU)=(f,L) on f1(U).i\Theta_{L,\varphi_{U}}\geqslant 0\quad\textrm{and}\quad\mathscr{I}(\varphi_{U})=\mathscr{I}(f,\|L\|)\textrm{ on }f^{-1}(U).

After replacing the φ\varphi before by {f1(U),φU}\{f^{-1}(U),\varphi_{U}\}, the isomorphism (3.9) is preserved.

Then we furthermore assume that there exists a section ss of some multiple LmL^{m} such that {s=0}Xo\{s=0\}\subseteq X^{o}. At this time (VZ)Xo=VXo(V\setminus Z)\cap X^{o}=V\cap X^{o}, and φU\varphi_{U} is smooth outside of XoX^{o}. As a consequence, we have

E(heφU)=E(h)(φU),E(h\cdot e^{-\varphi_{U}})=E(h)\otimes\mathscr{I}(\varphi_{U}),

and (3.9) is reformulated as

Proposition 3.3 ((Hodge-type isomorphism)).

Assume that there exists a section ss of some multiple LmL^{m} such that {s=0}Xo\{s=0\}\subseteq X^{o}. Then

Rqf(KXLE(h)(f,L))fn,q(LE(h)(f,L)).R^{q}f_{\ast}(K_{X}\otimes L\otimes E(h)\otimes\mathscr{I}(f,\|L\|))\simeq f_{\ast}\mathcal{H}^{n,q}(L\otimes E(h)\otimes\mathscr{I}(f,\|L\|)).

4 Injectivity theorem and vanishing theorem

We first prove Theorem 1.3, i.e. a Kollár-type injectivity and torsion-freeness theorem. Note that throughout the rest part of this paper, f:XYf:X\rightarrow Y is a proper, locally Kähler morphism from a complex manifold XX to a reduced, irreducible analytic space YY. Every connected component of XX is mapped surjectively to YY. Let ll be the dimension of a general fibre FF of ff. XoX^{o} is a Zariski open subset of XX, and (E,h)(E,h) is a tame, Nakano semi-positive vector bundle over XoX^{o}. Denote the fact that iΘE,hi\Theta_{E,h} is semi-positive in the sense of Nakano by iΘE,hNak0i\Theta_{E,h}\geqslant_{\textrm{Nak}}0.

4.1 A Kollár-type injectivity theorem

Theorem 4.1 ((=Theorem 1.3)).

Let LL be a holomorphic line bundle on XX with κ(L,f)0\kappa(L,f)\geqslant 0. For a (non-zero) section ss of some multiple Lm1L^{m-1} such that {s=0}Xo\{s=0\}\subseteq X^{o}, the multiplication map induced by the tensor product with ss

Φ:Rqf(KXLE(h)(f,L))Rqf(KXLmE(h)(f,Lm))\Phi:R^{q}f_{\ast}(K_{X}\otimes L\otimes E(h)\otimes\mathscr{I}(f,\|L\|))\rightarrow R^{q}f_{\ast}(K_{X}\otimes L^{m}\otimes E(h)\otimes\mathscr{I}(f,\|L^{m}\|))

is well-defined and injective for any q0q\geqslant 0. In particular, Rqf(KXLE(h)(f,L))R^{q}f_{\ast}(K_{X}\otimes L\otimes E(h)\otimes\mathscr{I}(f,\|L\|)) is torsion-free for every qq.

Proof.

Let {U}\{U\} be a Stein covering of YY. Let rUr_{U} be a smooth strictly plurisubharmonic exhaustion function on UU. From the discussion in Sect.2.1, there exits a collection of singular metrics

φ={f1(U),φU}\varphi=\{f^{-1}(U),\varphi_{U}\}

with algebraic singularities, such that

(f,L)=(φU),(f,Lm1)=((m1)φU) and (f,Lm)=(mφU) on f1(U).\mathscr{I}(f,\|L\|)=\mathscr{I}(\varphi_{U}),\mathscr{I}(f,\|L^{m-1}\|)=\mathscr{I}((m-1)\varphi_{U})\textrm{ and }\mathscr{I}(f,\|L^{m}\|)=\mathscr{I}(m\varphi_{U})\textrm{ on }f^{-1}(U).

Then in view of Proposition 3.3 it is left to prove that

s:fn,q(LE(h)(φ))fn,q(LmE(h)(mφ))\otimes s:f_{\ast}\mathcal{H}^{n,q}(L\otimes E(h)\otimes\mathscr{I}(\varphi))\rightarrow f_{\ast}\mathcal{H}^{n,q}(L^{m}\otimes E(h)\otimes\mathscr{I}(m\varphi))

is well-defined and injective.

Let αn,q(f1(U),LE(h)(φU),frU)\alpha\in\mathcal{H}^{n,q}(f^{-1}(U),L\otimes E(h)\otimes\mathscr{I}(\varphi_{U}),f^{\ast}r_{U}), and let ZZ be the closed subvariety such that φ\varphi is smooth on f1(U)Zf^{-1}(U)\setminus Z. Then ZXo=Z\cap X^{o}=\emptyset by assumption. By definition, there exists a sequence {αl}\{\alpha_{l}\} such that

αlφU,h,ωln,q(f1(U)Xo,LE),αl[α|f1(U)Xo],e(¯frU)αl=0\alpha_{l}\in\mathcal{H}^{n,q}_{\varphi_{U},h,\omega_{l}}(f^{-1}(U)\cap X^{o},L\otimes E),\alpha_{l}\in[\alpha|_{f^{-1}(U)\cap X^{o}}],e(\bar{\partial}f^{\ast}r_{U})^{\ast}\alpha_{l}=0

and limαl=α|f1(U)Xo\lim\alpha_{l}=\alpha|_{f^{-1}(U)\cap X^{o}} in the sense of L2L^{2}-topology. Apply formula (3.5) to

αl on f1(U)Xo,\alpha_{l}\textrm{ on }f^{-1}(U)\cap X^{o},

we obtain

0=¯αlφU,h,ωl2+¯φU,hαlφU,h,ωl2=φU,hαlφU,h,ωl2+(iΘLE,φU,hΛαl,αl)φU,h,ωl.\begin{split}0=&\|\bar{\partial}\alpha_{l}\|^{2}_{\varphi_{U},h,\omega_{l}}+\|\bar{\partial}^{\ast}_{\varphi_{U},h}\alpha_{l}\|^{2}_{\varphi_{U},h,\omega_{l}}\\ =&\|\partial^{\ast}_{\varphi_{U},h}\alpha_{l}\|^{2}_{\varphi_{U},h,\omega_{l}}+(i\Theta_{L\otimes E,\varphi_{U},h}\Lambda\alpha_{l},\alpha_{l})_{\varphi_{U},h,\omega_{l}}.\end{split}

Remember that iΘLE,φU,hNak0i\Theta_{L\otimes E,\varphi_{U},h}\geqslant_{\textrm{Nak}}0. Thus, φUαl=0\partial_{\varphi_{U}}^{\ast}\alpha_{l}=0 and

(iΘLE,φU,hΛαl,αl)φU,h,ωl=0.(i\Theta_{L\otimes E,\varphi_{U},h}\Lambda\alpha_{l},\alpha_{l})_{\varphi_{U},h,\omega_{l}}=0.

Now for an

sH0(X,Lm1),s\in H^{0}(X,L^{m-1}),

certainly we have ¯(sα)=0\bar{\partial}(s\alpha)=0 and e(¯frU)(sαl)=0e(\bar{\partial}f^{\ast}r_{U})^{\ast}(s\alpha_{l})=0. Let 𝔞(f,|L|)\mathfrak{a}(f,|L|) be the base-ideal of |L||L| relative to ff, so s𝔞(f,|Lm1|)s\in\mathfrak{a}(f,|L^{m-1}|). Then

[sα]Hq(f1(U),KXLmE(h)(mφU))[s\alpha]\in H^{q}(f^{-1}(U),K_{X}\otimes L^{m}\otimes E(h)\otimes\mathscr{I}(m\varphi_{U}))

by Proposition 2.1. Then due to Proposition 3.2, there exists a sequence {βl}\{\beta_{l}\} on f1(U)Xof^{-1}(U)\cap X^{o} such that ¯βl=¯mφU,hβl=e(¯frU)βl=0\bar{\partial}\beta_{l}=\bar{\partial}^{\ast}_{m\varphi_{U},h}\beta_{l}=e(\bar{\partial}f^{\ast}r_{U})^{\ast}\beta_{l}=0 and βl[(sα)|f1(U)Xo]\beta_{l}\in[(s\alpha)|_{f^{-1}(U)\cap X^{o}}]. It is left to prove that βl=sαl\beta_{l}=s\alpha_{l}. Indeed, since βl,sαl[(sα)|f1(U)Xo]\beta_{l},s\alpha_{l}\in[(s\alpha)|_{f^{-1}(U)\cap X^{o}}], there exits an LmEL^{m}\otimes E-valued (n,q1)(n,q-1)-form γl\gamma_{l} on f1(U)Xof^{-1}(U)\cap X^{o} such that sαl=βl+¯γls\alpha_{l}=\beta_{l}+\bar{\partial}\gamma_{l}. Now apply (3.1) and (3.2) to sαls\alpha_{l} on f1(U)Xof^{-1}(U)\cap X^{o} after shrinking UU if necessary, we obtain that

¯mφU,h(sαl)mφU,h,ωl2=mφU,h(sαl)mφU,h,ωl2+(iΘLmE,mφU,hΛ(sαl),sαl)mφU,h,ωl+[mφU,h(sαl),e(frU)(sαl)]mφU,h,ωl.\begin{split}\|\bar{\partial}^{\ast}_{m\varphi_{U},h}(s\alpha_{l})\|^{2}_{m\varphi_{U},h,\omega_{l}}=&\|\partial^{\ast}_{m\varphi_{U},h}(s\alpha_{l})\|^{2}_{m\varphi_{U},h,\omega_{l}}+(i\Theta_{L^{m}\otimes E,m\varphi_{U},h}\Lambda(s\alpha_{l}),s\alpha_{l})_{m\varphi_{U},h,\omega_{l}}\\ &+[\partial^{\ast}_{m\varphi_{U},h}(s\alpha_{l}),e(\partial f^{\ast}r_{U})^{\ast}(s\alpha_{l})]_{m\varphi_{U},h,\omega_{l}}.\end{split}

Note mφU,h(sαl)=¯(sαl)=s¯αl=sφU,hαl=0\partial^{\ast}_{m\varphi_{U},h}(s\alpha_{l})=-\ast\bar{\partial}\ast(s\alpha_{l})=-s\ast\bar{\partial}\ast\alpha_{l}=s\partial^{\ast}_{\varphi_{U},h}\alpha_{l}=0. On the other hand,

0(iΘLmE,mφU,hΛ(sαl),sαl)mφU,h,ωlsupX(|s|2e(m1)φU)(iΘLmE,φU,hΛαl,αl)φU,h,ωlmsupX(|s|2e(m1)φU)(iΘLE,φU,hΛαl,αl)φU,h,ωl=0.\begin{split}0\leqslant&(i\Theta_{L^{m}\otimes E,m\varphi_{U},h}\Lambda(s\alpha_{l}),s\alpha_{l})_{m\varphi_{U},h,\omega_{l}}\\ \leqslant&\sup_{X}(|s|^{2}e^{-(m-1)\varphi_{U}})(i\Theta_{L^{m}\otimes E,\varphi_{U},h}\Lambda\alpha_{l},\alpha_{l})_{\varphi_{U},h,\omega_{l}}\\ \leqslant&m\sup_{X}(|s|^{2}e^{-(m-1)\varphi_{U}})(i\Theta_{L\otimes E,\varphi_{U},h}\Lambda\alpha_{l},\alpha_{l})_{\varphi_{U},h,\omega_{l}}\\ =&0.\end{split}

Since supX(|s|2e(m1)φU)\sup_{X}(|s|^{2}e^{-(m-1)\varphi_{U}}) is obviously finite, we obtain that

(iΘLmE,mφU,hΛ(sαl),sαl)mφU,h,ωl=0.(i\Theta_{L^{m}\otimes E,m\varphi_{U},h}\Lambda(s\alpha_{l}),s\alpha_{l})_{m\varphi_{U},h,\omega_{l}}=0.

In summary,

¯mφU,h(sαl)mφU,h,ωl2=0.\|\bar{\partial}^{\ast}_{m\varphi_{U},h}(s\alpha_{l})\|^{2}_{m\varphi_{U},h,\omega_{l}}=0.

Then we have

¯mφU,h¯γlmφU,h,ωl2=¯mφU,h(sαlβl)mφU,h,ωl2=0.\begin{split}&\|\bar{\partial}^{\ast}_{m\varphi_{U},h}\bar{\partial}\gamma_{l}\|^{2}_{m\varphi_{U},h,\omega_{l}}\\ =&\|\bar{\partial}^{\ast}_{m\varphi_{U},h}(s\alpha_{l}-\beta_{l})\|^{2}_{m\varphi_{U},h,\omega_{l}}\\ =&0.\end{split}

In other words, ¯mφU,h¯γl=0\bar{\partial}^{\ast}_{m\varphi_{U},h}\bar{\partial}\gamma_{l}=0. Observe that e(¯frU)¯γl=e(¯frU)(sαlβl)=0e(\bar{\partial}f^{\ast}r_{U})^{\ast}\bar{\partial}\gamma_{l}=e(\bar{\partial}f^{\ast}r_{U})^{\ast}(s\alpha_{l}-\beta_{l})=0. Therefore by (3.1),

¯γlmφU,h,ωl2=(γl,¯mφU,h¯γl)mφU,h,ωl+[γl,e(¯frU)¯γl]mφU,h,ωl=0.\|\bar{\partial}\gamma_{l}\|^{2}_{m\varphi_{U},h,\omega_{l}}=(\gamma_{l},\bar{\partial}^{\ast}_{m\varphi_{U},h}\bar{\partial}\gamma_{l})_{m\varphi_{U},h,\omega_{l}}+[\gamma_{l},e(\bar{\partial}f^{\ast}r_{U})^{\ast}\bar{\partial}\gamma_{l}]_{m\varphi_{U},h,\omega_{l}}=0.

We conclude that ¯γl=0\bar{\partial}\gamma_{l}=0. Equivalently, βl=sαl\beta_{l}=s\alpha_{l} on f1(U)Xof^{-1}(U)\cap X^{o}. Now

lim(sα)|f1(U)XoβlmφU,h,ω2=lim(sα)|f1(U)XosαlmφU,h,ω2supX(|s|2e(m1)φU)limα|f1(U)XoαlφU,h,ω2=0.\begin{split}\lim\|(s\alpha)|_{f^{-1}(U)\cap X^{o}}-\beta_{l}\|^{2}_{m\varphi_{U},h,\omega}&=\lim\|(s\alpha)|_{f^{-1}(U)\cap X^{o}}-s\alpha_{l}\|^{2}_{m\varphi_{U},h,\omega}\\ &\leqslant\sup_{X}(|s|^{2}e^{-(m-1)\varphi_{U}})\lim\|\alpha|_{f^{-1}(U)\cap X^{o}}-\alpha_{l}\|^{2}_{\varphi_{U},h,\omega}\\ &=0.\end{split}

In summary,

sαn,q(f1(U),LmE(h)(mφ),frU).s\alpha\in\mathcal{H}^{n,q}(f^{-1}(U),L^{m}\otimes E(h)\otimes\mathscr{I}(m\varphi),f^{\ast}r_{U}).

Then we have successfully proved that

s:n,q(f1(U),LE(h)(φU),frU)n,q(f1(U),LmE(h)(mφU),frU)\otimes s:\mathcal{H}^{n,q}(f^{-1}(U),L\otimes E(h)\otimes\mathscr{I}(\varphi_{U}),f^{\ast}r_{U})\rightarrow\mathcal{H}^{n,q}(f^{-1}(U),L^{m}\otimes E(h)\otimes\mathscr{I}(m\varphi_{U}),f^{\ast}r_{U})

is well-defined. It is similar for a general Stein subset VUV\subseteq U. As a result, the sheaf morphism

s:fn,q(LE(h)(φ))fn,q(LmE(h)(mφ))\otimes s:f_{\ast}\mathcal{H}^{n,q}(L\otimes E(h)\otimes\mathscr{I}(\varphi))\rightarrow f_{\ast}\mathcal{H}^{n,q}(L^{m}\otimes E(h)\otimes\mathscr{I}(m\varphi))

is also well-defined. The injectivity is then obvious.

Observe that if we substitute ss by an arbitrary holomorphic function gg on XX, everything is still going smoothly. Thus we have the following conclusion: the map induced by multiplied with gg:

×g:Rqf(KXLE(h)(f,L))Rqf(KXLE(h)(f,L))\times g:R^{q}f_{\ast}(K_{X}\otimes L\otimes E(h)\otimes\mathscr{I}(f,\|L\|))\rightarrow R^{q}f_{\ast}(K_{X}\otimes L\otimes E(h)\otimes\mathscr{I}(f,\|L\|))

is well-defined and injective. This immediately implies that Rqf(KXLE(h)(f,L))R^{q}f_{\ast}(K_{X}\otimes L\otimes E(h)\otimes\mathscr{I}(f,\|L\|)) is torsion-free. ∎

4.2 A relative Nadel-type vanishing theorem

Firstly, let’s recall an absolute version of the Nadel-type vanishing theorem, which is a simple variant of [Wu22], Theorem 1.3.

Theorem 4.2.

Let XX be a compact Kähler manifold of dimension nn, and let LL be a pseudo-effective line bundle on XX. Let XoX^{o} be a Zariski open subset of XX, and let (E,h)(E,h) be a tame, Nakano semi-positive vector bundle of rank rr over XoX^{o}. Assume that there exists a section ss of some multiple LmL^{m} such that {s=0}Xo\{s=0\}\subseteq X^{o}. Then

Hq(X,KXLE(h)(L))=0H^{q}(X,K_{X}\otimes L\otimes E(h)\otimes\mathscr{I}(\|L\|))=0

for q>nκ(L)q>n-\kappa(L).

(Sketch of Proof).

The proof is almost the same as [Wu22] except minor adjustment, so we only outline the main stream here.

We first use a log-resolution μ:X~X\mu:\tilde{X}\rightarrow X to reduce the vanishing as

Hq(X~,KX~L^μE(h)),H^{q}(\tilde{X},K_{\tilde{X}}\otimes\hat{L}\otimes\mu^{\ast}E(h)),

where L^=μL𝒪X~(λipEi)\hat{L}=\mu^{\ast}L\otimes\mathcal{O}_{\tilde{X}}(-\sum\lfloor\frac{\lambda_{i}}{p}\rfloor E_{i}). Note EiE_{i} is the prime component of the exceptional divisor with certain coefficients λip\frac{\lambda_{i}}{p}. Let eie_{i} be the defining section of EiE_{i}.

Denote L~=μL𝒪X~(λipEi)\tilde{L}=\mu^{\ast}L\otimes\mathcal{O}_{\tilde{X}}(-\sum\frac{\lambda_{i}}{p}E_{i}). The crucial observation is that we can endow L~p\tilde{L}^{p}, which is a \mathbb{Z}-bundle, a smooth metric ψ\psi with semi-positive associated curvature, and κ(L~p)=κ(L)\kappa(\tilde{L}^{p})=\kappa(L). Now we solve the complex Monge–Ampère equations

iΘL~p,φε+εω>0and(iΘL~p,φε+εω)n=Cεωni\Theta_{\tilde{L}^{p},\varphi_{\varepsilon}}+\varepsilon\omega>0\quad\textrm{and}\quad(i\Theta_{\tilde{L}^{p},\varphi_{\varepsilon}}+\varepsilon\omega)^{n}=C_{\varepsilon}\omega^{n}

for every ε>0\varepsilon>0, to obtain smooth metrics {φε}\{\varphi_{\varepsilon}\} on L~p\tilde{L}^{p}.

We endow L^\hat{L} with singular metric

ϕε=1p(δφε+(1δ)ψ)+{λip}log|ei|2,\phi_{\varepsilon}=\frac{1}{p}(\delta\varphi_{\varepsilon}+(1-\delta)\psi)+\sum\{\frac{\lambda_{i}}{p}\}\log|e_{i}|^{2},

where δ>0\delta>0 is a sufficiently small number which will be fixed later. Denote by 0<a1an0<a_{1}\leqslant\cdots\leqslant a_{n} and 0<a^1a^n0<\hat{a}_{1}\leqslant\cdots\leqslant\hat{a}_{n}, respectively, the eigenvalues of the curvature forms iΘL~p,φε+εωi\Theta_{\tilde{L}^{p},\varphi_{\varepsilon}}+\varepsilon\omega and iΘL^,ϕε+2εpωi\Theta_{\hat{L},\phi_{\varepsilon}}+\frac{2\varepsilon}{p}\omega at every point xX~x\in\tilde{X}, with respect to the base Kähler metric ω(x)\omega(x). We apply (3.2) on XoX^{o} for every L^μE\hat{L}\otimes\mu^{\ast}E-valued (n,q)(n,q)-form α\alpha to obtain

¯αϕε,h2+¯αϕε,h2=ϕε,hαϕε,h2+(iΘL^μE,ϕε,hΛα,α)ϕε,hX~r(a^1++a^q2qεp)|α|ϕε,h2𝑑Vω.\begin{split}\|\bar{\partial}\alpha\|^{2}_{\phi_{\varepsilon},h}+\|\bar{\partial}^{\ast}\alpha\|^{2}_{\phi_{\varepsilon},h}&=\|\partial^{\ast}_{\phi_{\varepsilon,h}}\alpha\|^{2}_{\phi_{\varepsilon},h}+(i\Theta_{\hat{L}\otimes\mu^{\ast}E,\phi_{\varepsilon},h}\Lambda\alpha,\alpha)_{\phi_{\varepsilon},h}\\ &\geqslant\int_{\tilde{X}}r(\hat{a}_{1}+\cdots+\hat{a}_{q}-\frac{2q\varepsilon}{p})|\alpha|^{2}_{\phi_{\varepsilon},h}dV_{\omega}.\end{split} (4.1)

Return to the proof of vanishing. Let us take a cohomology class

[β]Hq(X~,KX~L^μE(h)).[\beta]\in H^{q}(\tilde{X},K_{\tilde{X}}\otimes\hat{L}\otimes\mu^{\ast}E(h)).

By using the de Rham–Weil isomorphism, we obtain a representative β\beta which is a smooth L^μE\hat{L}\otimes\mu^{\ast}E-valued (n,q)(n,q)-form.

Then we use the canonical L2L^{2}-estimate [Dem12] against ϕε,h\phi_{\varepsilon},h to obtain elements uε,vεu_{\varepsilon},v_{\varepsilon} such that β=uε+¯vε\beta=u_{\varepsilon}+\bar{\partial}v_{\varepsilon}. Moreover, (4.1) implies that

vεϕε,h22qεpX~1a^1++a^q|β|ϕε,h2𝑑Vω.\|v_{\varepsilon}\|^{2}_{\phi_{\varepsilon},h}\leqslant\frac{2q\varepsilon}{p}\int_{\tilde{X}}\frac{1}{\hat{a}_{1}+\cdots+\hat{a}_{q}}|\beta|^{2}_{\phi_{\varepsilon},h}dV_{\omega}. (4.2)

Denote γε:=qεp(a^1++a^q)\gamma_{\varepsilon}:=\frac{q\varepsilon}{p(\hat{a}_{1}+\cdots+\hat{a}_{q})}, then

γεmin(1,Cδ1ε1(nκ(L))/q(aq+1an)1/q)\gamma_{\varepsilon}\leqslant\min(1,C\delta^{-1}\varepsilon^{1-(n-\kappa(L))/q}(a_{q+1}\cdots a_{n})^{1/q})

for a universal constant CC. Here we use the fact that

Cε=X~(c1(L~p)+εω)nX~ωnCεnκ(L)C_{\varepsilon}=\frac{\int_{\tilde{X}}(c_{1}(\tilde{L}^{p})+\varepsilon\omega)^{n}}{\int_{\tilde{X}}\omega^{n}}\geqslant C^{\prime}\varepsilon^{n-\kappa(L)}

for a universal constant CC^{\prime}. Notice that

X~aq+1an𝑑VωX~(iΘL~p,φε+εω)nqωq=(c1(L~p)+ε[ω])nq[ω]qC′′,\int_{\tilde{X}}a_{q+1}\cdots a_{n}dV_{\omega}\leqslant\int_{\tilde{X}}(i\Theta_{\tilde{L}^{p},\varphi_{\varepsilon}}+\varepsilon\omega)^{n-q}\wedge\omega^{q}=(c_{1}(\tilde{L}^{p})+\varepsilon[\omega])^{n-q}[\omega]^{q}\leqslant C^{\prime\prime},

hence the functions (aq+1an)1/q(a_{q+1}\cdots a_{n})^{1/q} are uniformly bounded in L1L^{1}-norm as ε\varepsilon tens to zero. So γε\gamma_{\varepsilon} converges almost everywhere to zero as ε\varepsilon tends to zero when q>nκ(L)q>n-\kappa(L).

In the end, some standard analysis shows that for a small enough δ\delta (independent of ε\varepsilon),

limε02qεpX~1a^1++a^q|β|ϕε,h2𝑑Vω=limε02X~γε|β|ϕε,h2𝑑Vω=0.\lim_{\varepsilon\rightarrow 0}\frac{2q\varepsilon}{p}\int_{\tilde{X}}\frac{1}{\hat{a}_{1}+\cdots+\hat{a}_{q}}|\beta|^{2}_{\phi_{\varepsilon},h}dV_{\omega}=\lim_{\varepsilon\rightarrow 0}2\int_{\tilde{X}}\gamma_{\varepsilon}|\beta|^{2}_{\phi_{\varepsilon},h}dV_{\omega}=0.

Therefore vεv_{\varepsilon} converges to zero as ε\varepsilon tends to zero. Equivalently, β\beta is actually a boundary, and the desired vanishing is proved. ∎

Now we should prove Theorem 1.2, i.e. a relative Nadel-type vanishing theorem.

Theorem 4.3 ((=Theorem 1.2)).

Let LL be a holomorphic line bundle with κ(L,f)0\kappa(L,f)\geqslant 0. Assume that there exists a section ss of some multiple LmL^{m} such that {s=0}Xo\{s=0\}\subseteq X^{o}. Then

Rqf(KXLE(h)(f,L))=0R^{q}f_{\ast}(K_{X}\otimes L\otimes E(h)\otimes\mathscr{I}(f,\|L\|))=0

for q>lκ(L,f)q>l-\kappa(L,f).

Proof.

Let ZZ be the set of the critical value of ff, which is a closed subvariety of YY. As a result, XyX_{y} is a compact Kähler manifold when yf(Xo)Zy\in f(X^{o})\setminus Z. Then we apply Theorem 4.2 to obtain that

Hq(Xy,KXyL|XyE(h)|Xy(L|Xy))=0H^{q}(X_{y},K_{X_{y}}\otimes L|_{X_{y}}\otimes E(h)|_{X_{y}}\otimes\mathscr{I}(\|L|_{X_{y}}\|))=0

for q>lκ(L|Xy)q>l-\kappa(L|_{X_{y}}).

Now we claim that there exists a subset VV of YY with zero Lebesgue measure, such that κ(L|Xy)=κ(L,f)\kappa(L|_{X_{y}})=\kappa(L,f) and

(L|Xy)=(f,L)\mathscr{I}(\|L|_{X_{y}}\|)=\mathscr{I}(f,\|L\|)

when yYVy\in Y\setminus V. In fact, since f(Lm)f_{\ast}(L^{m}) is torsion-free for all mm, there exists a closed subvariety ZmZ_{m} such that f(Lm)f_{\ast}(L^{m}) is locally free on YZmY\setminus Z_{m}. Let V=(m=1Zm)ZV=(\cup^{\infty}_{m=1}Z_{m})\cup Z, which has zero Lebesgue measure. Then every section in H0(Xy,Lm)H^{0}(X_{y},L^{m}) extends to the whole XX and κ(L|Xy)=κ(L,f)\kappa(L|_{X_{y}})=\kappa(L,f) when yYVy\in Y\setminus V. Hence (L|Xy)=(f,L)\mathscr{I}(\|L|_{X_{y}}\|)=\mathscr{I}(f,\|L\|) by definition.

Now on f(Xo)Vf(X^{o})\setminus V we obtain

Rqf(KXLE(h)(f,L))=0R^{q}f_{\ast}(K_{X}\otimes L\otimes E(h)\otimes\mathscr{I}(f,\|L\|))=0

for q>lκ(L,f)q>l-\kappa(L,f). We then conclude that this vanishing actually holds on the whole YY since Rqf(KXLE(h)(f,L))R^{q}f_{\ast}(K_{X}\otimes L\otimes E(h)\otimes\mathscr{I}(f,\|L\|)) is torsion-free by Theorem 4.1. ∎

5 Applications

In this section we should discuss the applications on harmonic bundles. Firstly, combining Theorem 1.2 and Proposition 2.2, we obtain that

Corollary 5.1 ((=Corollary 1.2)).

Let LL be a holomorphic line bundle with κ(L,f)0\kappa(L,f)\geqslant 0. Assume that there exists a section ss of some multiple LmL^{m} such that {s=0}Xo\{s=0\}\subseteq X^{o}. Let (H,θ,h)(H,\theta,h) be a tame harmonic bundle over XoX^{o}, and let EE be a subbundle of HH with vanishing second fundamental form and θ¯(E)=0\bar{\theta}(E)=0. Then

Rqf(KXLE(h)(f,L))=0R^{q}f_{\ast}(K_{X}\otimes L\otimes E(h)\otimes\mathscr{I}(f,\|L\|))=0

for q>lκ(L,f)q>l-\kappa(L,f).

In particular, Saito’s SS-sheaf can be represented as such an E(h)E(h). The original construction of S-sheaves for a real variation of polarized Hodge structure

(𝕍,,,S)(\mathbb{V},\nabla,\mathcal{F}^{\cdot},S)

is based on the theory of Hodge module. We recommend [Sai91] as the reference. However, the original definition of S-sheaves is not involved in our paper. Instead, let’s recall the following equivalent description provided in [ScY20].

Proposition 5.1 ((c.f. [ScY20], Theorem A)).

Let (𝕍,,,S)(\mathbb{V},\nabla,\mathcal{F}^{\cdot},S) be a real variation of polarized Hodge structure on XoX^{o}. Denote by S(𝕍)=pmaxS(\mathbb{V})=\mathcal{F}^{p_{\max}} where pmax=max{p;p0}p_{\max}=\max\{p;\mathcal{F}^{p}\neq 0\}. Denote by hh the Hermitian metric on S(𝕍)S(\mathbb{V}) induced by the polarization SS. Denote by ICX(𝕍)\mathrm{IC}_{X}(\mathbb{V}) the intermediate extension of 𝕍\mathbb{V} on XX as a pure Hodge module and by S(ICX(𝕍))S(\mathrm{IC}_{X}(\mathbb{V})) the Saito’s S-sheaf associated to ICX(𝕍)\mathrm{IC}_{X}(\mathbb{V}). Then

KXS(𝕍)(h)S(ICX(𝕍)).K_{X}\otimes S(\mathbb{V})(h)\simeq S(\mathrm{IC}_{X}(\mathbb{V})).

Based on this result, we obtain that

Corollary 5.2 ((=Corollary 1.3)).

Let LL be a holomorphic line bundle on XX with κ(L,f)0\kappa(L,f)\geqslant 0. Assume that there exists a section ss of some multiple LmL^{m} such that {s=0}Xo\{s=0\}\subseteq X^{o}. Then

Rqf(S(ICX(𝕍))L(f,L))=0R^{q}f_{\ast}(S(\mathrm{IC}_{X}(\mathbb{V}))\otimes L\otimes\mathscr{I}(f,\|L\|))=0

for q>lκ(L,f)q>l-\kappa(L,f).

Proof.

Let (H,θ,hH)(H,\theta,h_{H}) be the tame harmonic bundle associated with (𝕍,,,S)(\mathbb{V},\nabla,\mathcal{F}^{\cdot},S). In order to apply Theorem 1.2, in the view of Proposition 2.3 and Proposition 5.1, it is enough to show that (S(𝕍),h)(S(\mathbb{V}),h) is Nakano semi-positive. There are two methods available.

In the view of Proposition 2.2, we could show that θ¯(S(𝕍))=0\bar{\theta}(S(\mathbb{V}))=0. However, it is obvious for the reason of degree. Alternatively, the Nakano semi-positivity is the direct consequence of Schmid’s curvature calculation. (See [Sch73], Lemma 7.18.) ∎

References

  • [Cao14] J. Cao, Numerical dimension and a Kawamata–Viehweg–Nadel-type vanishing theorem on compact Kähler manifolds, Compos. Math. 60 (2002), 295-313.
  • [CKS86] E. Cattani, A. Kaplan, W. Schmid, Degeneration of Hodge Structures, Ann. of Math. 123 (1986), 457-535.
  • [Dem82] J.-P. Demailly, Estimations L2L^{2} pour l’opérateur ¯\bar{\partial} d’un fibré vectoriel holomorphe semi-positif au-dessus d’une variété kählérienne complète, Ann.  Sci.  École  Norm.  Sup.  (4) 15 (1982), 457-511.
  • [Dem93] J.-P. Demailly,A numerical criterion for very ample line bundles, J.  Diff.  Geom. 37 (1993), 323-374.
  • [Dem12] J.-P. Demailly, Analytic methods in algebraic geometry, Surveys of Modern Mathematics 1, International Press, Somerville, MA; Higher Education Press, Beijing, 2012.
  • [DoX84] H. Donnelly, F. Xavier, On the differential forms spectrum of negatively curved Riemann manifolds, Am.  J.  Math. 106 (1984), 169-186.
  • [Eno93] I. Enoki, Kawamata–Viehweg vanishing theorem for compact Kähler manifolds, Einstein metrics and Yang–Mills connections. Marcel Dekker, (1993), 59-68.
  • [Fuj12] O. Fujino, A transcendental approach to Kollár’s injectivity theorem, Osaka  J.  Math. 49 (2012), 833-852.
  • [Har77] R. Hartshorne, Algebraic geometry, Graduate Texts in Mathematics 52, Springer-Verlag, New York-Heidelberg, 1977.
  • [Ko86a] J. Kollár, Higher direct images of dualizing sheaves. I, Ann.  of  Math.  (2) 123 (1986), 11-42.
  • [Ko86b] J. Kollár, Higher direct images of dualizing sheaves. II, Ann.  of  Math.  (2) 124 (1986), 171-202.
  • [Laz04a] R. Lazarsfeld, Positivity in algebraic geometry. I. Classical setting: line bundles and linear series, Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics, Springer-Verlag, Berlin, 2004. xviii+387 pp. ISBN: 3-540-22533-1.
  • [Laz04b] R. Lazarsfeld, Positivity in algebraic geometry. II. Positivity for vector bundles, and multiplier ideals, Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics, Springer-Verlag, Berlin, 2004. xviii+385 pp. ISBN: 3-540-22534-X.
  • [Mat14] S. Matsumura, A Nadel vanishing theorem via injective theorems, Math.  Ann. 359 (2014), 785-802.
  • [Mat15a] S. Matsumura, Injectivity theorems with multiplier ideal sheaves and their applications, Complex analysis and geometry, 241-255. Springer Proc. Math. Stat., 144, 2015.
  • [Mat15b] S. Matsumura, A Nadel vanishing theorem for metrics with minimal singularities on big line bundles, Adv.  Math. 280 (2015), 188-207.
  • [Mat18] S. Matsumura, An injectivity theorem with multiplier ideal sheaves of singular metrics with transcendental singularities, J.  Algebraic Geom. 27 (2018), 305-337.
  • [Nad90] Alan M. Nadel, Multiplier ideal sheaves and Kähler–Einstein metrics of positive scalar curvature, Ann.  of  Math.  (2) 132 (1990), 549-596.
  • [Ohs02] T. Ohsawa, Analysis of several complex variables, Translations of Mathematical Monographs. 211, AMS, 2002.
  • [Sai91] M. Saito, On Kollár’s conjecture, Several complex variables and complex geometry, Part 2 (1989), 509-517.
  • [Sch73] W. Schmid, Variation of Hodge structure: the singularities of the period mapping, Invent.  Math. 22 (1973), 211-319.
  • [ScY20] C. Schnell, R. Yang, Hodge modules and singular hermitian metrics, arXiv:2003.09064.
  • [ShZ21a] J. Shentu, C. Zhao, L2L^{2}-Extension of Adjoint bundles and Kollár’s Conjecture, Math.  Ann.  (published on line).
  • [ShZ21b] J. Shentu, C. Zhao, L2L^{2}-Dolbeault resolutions of the lowest Hodge piece of a Hodge module and an application to the relative Fujita conjecture, arXiv:2104.04905.
  • [ShZ22] J. Shentu, C. Zhao, Kollár’s package for twisted Saito’s S-sheaves, arXiv:2210.04131.
  • [Sim92] C. Simpson, Higgs bundles and local systems, Inst.  Hautes  Études  Sci.  Publ.  Math. 75 (1992), 5-95.
  • [Tak95] K. Takegoshi, Higher direct images of canonical sheaves tensorized with semi-positive vector bundles by proper Kähler morphisms, Math.  Ann. 303 (1995), 389-416.
  • [Wu17] J. Wu, A Kollár-type vanishing theorem, Math. Z. 295 (2020), 331-340.
  • [Wu20] J. Wu, An eigenvalue estimate for the ¯\bar{\partial}-Laplacian associated to a nef line bundle, arXiv:2003.05187.
  • [Wu22] J. Wu, The Kawamata–Viehweg–Nadel-type vanishing theorem and the asymptotic multiplier ideal sheaf, Math.  Z.  302 (2022), 1393-1407.